You are on page 1of 91

Tsunami evacuation: Lessons from

the Great East Japan earthquake and


tsunami of March 11th 2011
S. Fraser G.S. Leonard
I. Matsuo H. Murakami

GNS Science Report 2012/17


April 2012
BIBLIOGRAPHIC REFERENCE

Fraser, S.; Leonard, G.S.; Matsuo, I. and Murakami, H. 2012. Tsunami


evacuation: Lessons from the Great East Japan earthquake and tsunami of
March 11th 2011, GNS Science Report 2012/17. 89 p.

S. Fraser, GNS Science/Massey University Joint Centre for Disaster Research, PO Box 756,
Wellington, New Zealand

G.S. Leonard, GNS Science, PO Box 30368, Lower Hutt

I. Matsuo, Crisis and Environment Management Policy Institute (CeMI),


1-22-505 Wakaba, Shinjyuku-ku, Tokyo 160-0011, Japan.

H Murakami, Division of Environmental Science and Engineering, Yamaguchi University,


Tokiwadai, Ube 755-8611, Japan.

© Institute of Geological and Nuclear Sciences Limited, 2012


ISSN 1177-2425
ISBN 978-0-478-19897-3

i
2012

CONTENTS
ABSTRACT............................................................................................................................VII
KEYWORDS .........................................................................................................................VIII
1.0 INTRODUCTION .......................................................................................................... 1
1.1 Background to the event ................................................................................... 1
1.2 Goals, methodology and report structure ......................................................... 3
1.3 Interviews .......................................................................................................... 4
1.4 Tsunami preparedness in the T hoku region of Japan .................................... 6
1.4.1 Previous events and hazard estimation ................................................ 6
1.4.2 Hazard estimation and mapping ........................................................... 6
1.4.3 Education and evacuation exercises .................................................... 7
1.4.4 Earthquake early warning system and rapid official tsunami warnings . 7
1.4.5 Coastal defences .................................................................................. 8
1.5 Tsunami preparedness in the State of Washington and New Zealand ............. 8
1.5.1 Washington ........................................................................................... 8
1.5.2 New Zealand ....................................................................................... 11
2.0 INTERVIEW OUTCOMES .......................................................................................... 13
2.1 Community involvement in evacuation planning and preparedness ............... 13
2.2 Hazard/evacuation maps and signs ................................................................ 14
2.2.1 Hazard/evacuation maps .................................................................... 14
2.2.2 Hazard/evacuation signs ..................................................................... 20
2.3 Response to natural and informal tsunami warnings ...................................... 25
2.3.1 Natural warnings ................................................................................. 25
2.3.2 Informal warnings ................................................................................ 27
2.4 Formal warning messages and systems ........................................................ 28
2.4.1 Warning messages ............................................................................. 28
2.4.2 Warning systems ................................................................................ 29
2.5 Movements during and after evacuation ......................................................... 30
2.5.1 Delayed evacuation or non-evacuation ............................................... 30
2.5.2 Use of motor vehicles during evacuation ............................................ 32
2.5.3 Use of boats during evacuation .......................................................... 34
2.5.4 Schools evacuation ............................................................................. 34
2.5.5 Necessary relocation from evacuation refuges ................................... 35
2.5.6 Leaving evacuation refuges too early ................................................. 35
2.6 Impacts of coastal defence on warning response and evacuation ................. 36
2.7 Vertical evacuation buildings .......................................................................... 37
2.7.1 Building specification and levels of damage ....................................... 38
2.7.2 Observed examples of vertical evacuation buildings .......................... 38
2.7.3 Community engagement ..................................................................... 51
2.7.4 Evacuation building signage ............................................................... 51
2.7.5 Owner agreement ............................................................................... 52
2.7.6 24-hour access ................................................................................... 52
2.7.7 Evacuee Welfare ................................................................................. 53
3.0 RECOMMENDATIONS .............................................................................................. 54
3.1 Community involvement in evacuation planning and preparedness ............... 54
3.2 Hazard/evacuation maps and signs ................................................................ 54
3.2.1 Maps ................................................................................................... 54
3.2.2 Hazard/evacuation signs ..................................................................... 55
3.3 Response to natural and informal tsunami warnings ...................................... 55
3.4 Formal warning messages and systems ........................................................ 56
3.4.1 Warning messages ............................................................................. 56
3.4.2 Warning systems ................................................................................ 57

GNS Science Report 2012/17 ii


2012

3.5 Movements during and after evacuation ......................................................... 57


3.5.1 Delayed evacuation or non-evacuation ............................................... 57
3.5.2 Use of motor vehicles during evacuation ............................................ 57
3.5.3 Use of boats during evacuation .......................................................... 58
3.5.4 Schools evacuation ............................................................................. 58
3.5.5 Necessary relocation from evacuation refuges ................................... 59
3.5.6 Leaving evacuation refuges too early ................................................. 59
3.6 Impacts of coastal defence on warning response and evacuation ................. 59
3.7 Vertical evacuation buildings .......................................................................... 60
3.7.1 Building specification and levels of damage ....................................... 60
3.7.2 Community engagement ..................................................................... 60
3.7.3 Evacuation building signage ............................................................... 61
3.7.4 Owner agreement ............................................................................... 61
3.7.5 24-hour access ................................................................................... 61
3.7.6 Evacuee welfare ................................................................................. 61
3.8 Additional recommendations from observations and interviews ..................... 62
4.0 ACKNOWLEDGEMENTS .......................................................................................... 62
5.0 REFERENCES ........................................................................................................... 62

FIGURES
Figure 1 Map of locations investigated in this report, with coastal topography and epicentre of March
11th main shock indicated. Note the grey lines delineate prefectures. All of the locations
investigated lie within Iwate and Miyagi prefectures (labelled), which received the largest
tsunami wave heights and most of the casualties. Base map credits: GEBCO, NOAA,
National Geographic, DeLorme, Esri.................................................................................................. 5
Figure 2 Excerpt from the tsunami evacuation map Clallam Bay, Clallam County, Washington. This
map uses 2 zones, local sites of importance (fire department, etc.), evacuation routes and
assembly areas (Clallam County Emergency Management, n d). ................................................... 10
Figure 3 An example evacuation map in Washington. This map is for Long Beach and Ilwaco, Pacific
County, and illustrates the use of 3 zones, evacuation routes and assembly areas (Pacific
County Emergency Management, 2007).......................................................................................... 10
Figure 4 ‘Tsunami Hazard Zone’ sign with instructions to evacuate to high ground or inland in
response to an earthquake. ............................................................................................................. 11
Figure 5 ‘Tsunami Evacuation Route’ sign with arrow depicting the direction of evacuation. These
signs are placed regularly along roads in tsunami hazard zones. .................................................... 11
Figure 6 An example tsunami hazard map from Whananaki, New Zealand. .................................................. 12
Figure 7 An example tsunami evacuation map and information board as displayed outdoors (1.2 m
wide) in the Island Bay community, Wellington, New Zealand. ........................................................ 12
Figure 8 Extract from Tar tsunami hazard map, indicating three sections of tsunami wall (red (1958
section), black and yellow thick lines), tsunami gates (blue triangles and circles). The
inundation extent of the 1986 Meiji (thin red line) and 1933 Sh wa tsunami (thin blue line).
Evacuation refuges are marked as red triangles and welfare centres as red circles. Source:
Tar Town Disaster Prevention Office. ............................................................................................16
Figure 9 Kamaishi City evacuation map (Source: Kamaishi City Disaster Prevention Office). ....................... 16
Figure 10 funato City evacuation map (Source: funato City Disaster Management). ................................. 17
Figure 11 Excerpt from Kesennuma City evacuation map (Source: Kesennuma General Affairs
Department). .................................................................................................................................... 18
Figure 12 Minami-Sanriku hazard map with evacuation refuge locations indicated by numbered red
circles (Miyagi Prefectural Government, 2004). ...............................................................................19
Figure 13 Excerpt from the Ishinomaki tsunami hazard map with evacuation refuges (Ishinomaki City,
n d). .................................................................................................................................................. 19
Figure 14 Natori City hazard map (Source: Natori City Disaster Prevention Office)......................................... 20
Figure 15 A sign displayed above the road indicating to drivers that they are entering an estimated
tsunami inundation area. These signs were seen on roads within Kamaishi district. ....................... 20
Figure 16 A sign displayed above the road indicating to drivers that they are exiting an estimated
tsunami inundation area. These signs were seen on roads within Kamaishi district. ....................... 20
Figure 17 Tsunami evacuation route direction and distance sign displayed on a sign in
Hitachihamacho, Miyako City. .......................................................................................................... 21

GNS Science Report 2012/17 iii


2012

Figure 18 Tsunami evacuation route direction and distance sign displayed on a lamp-post in
Hitachihamacho, Miyako City. .......................................................................................................... 21
Figure 19 Tsunami evacuation route information board and directions (with distance to safe location) in
Matsushima Town, Miyagi Prefecture. ............................................................................................. 21
Figure 20 Tsunami evacuation route sign with direction and distance to safe location painted on the
pavement in Matsushima Town, Miyagi Prefecture. ......................................................................... 21
Figure 21 A sign in Matsushima Town, displaying the tsunami evacuation route. ........................................... 22
Figure 22 Tsunami evacuation route sign with solar powered lamp, on a road leading to high ground
out of Tar Town. ............................................................................................................................. 22
Figure 23 Evacuation route sign in funato City, indicating the distance to a “tsunami emergency place
of refuge”. Photo courtesy of EEFIT. ................................................................................................ 22
Figure 24 A sign in the centre of Matsushima Town, advising evacuation on experiencing an
earthquake. ...................................................................................................................................... 23
Figure 25 A sign indicating tsunami hazard, on a road out of Kamaishi City. Translation reads:
“Tsunami Attention! Tsunami inundation hazard zone”. ................................................................... 23
Figure 26 Tsunami evacuation muster point for children in Hitachihamacho, Miyako City. .............................. 23
Figure 27 Close-up view of tsunami evacuation muster point for children in Hitachihamacho, Miyako
City. Translation reads: “Tsunami shelter (Kuwagasaki elementary school)”. .................................. 23
Figure 28 A marker post indicating past inundation height in central Minami-Sanriku Town. The
Japanese text gives notification that the Shizugawa hospital is a designated vertical
evacuation building. The Japanese text informs people to evacuate to designated
evacuation places such as parks, highlands, and Shizugawa elementary school and
Shizugawa hospital. ......................................................................................................................... 24
Figure 29 A marker post indicating past inundation height in central Minami-Sanriku Town. This marker
was found close to a care home (1 km inland), and displays the standard icon representing
evacuation to high ground. ............................................................................................................... 24
Figure 30 Tsunami height marker on a building wall in funato City, indicating the maximum
inundation height at this point during the Chile tsunami on May 24th 1960. Photograph
courtesy of EEFIT. ........................................................................................................................... 24
Figure 31 Landslide hazard and evacuation map on an information board in Ishinomaki City. ........................ 25
Figure 32 A wireless receiver shown by civil protection officials in Minami-Sanriku. These units are
rented by the local government to every household in the town, and resulted in effective
warning dissemination on March 11th 2011. ..................................................................................... 30
Figure 33 Building damage and fatality rate for seven areas of Yuriage, Natori City. Source: (Murakami
and Kashiwabara, 2011). ................................................................................................................. 32
Figure 34 Designated evacuation locations in Kamaishi City. An estimated 50 people evacuated to this
apartment block (A) and government offices (B). The entire area shown was inundated;
inundation depth was 8 m in the vicinity of building (A). ................................................................... 39
Figure 35 Apartment block in Kamaishi City, which was inundated to the 3rd storey........................................ 39
Figure 36 Government offices in Kamaishi City. This building is raised several metres above sea level
st
and sustained damage to only the 1 storey windows ..................................................................... 39
Figure 37 The Hotel Horaikan at Unosumai, Kamaishi City, with official signage displayed at the 3rd
nd
floor and very little external damage. The hotel was inundated to the ceiling of the 2 storey ........ 40
Figure 38 Locations of 5 designated vertical evacuation location visited in the downtown area of
funato City: (A) Plaza Hotel; (B) Yasuka office/apartment buildings; (C) Maiya Shopping
Centre and car park. ........................................................................................................................ 40
Figure 39 Maiya shopping centre, which was inundated to the 3rd storey ........................................................ 41
Figure 40 Maiya multi-storey car park showing damaged railing at the 2nd storey. .......................................... 41
Figure 41 Yasuka private commercial and apartment block number 1 shows very little exterior
damage. ........................................................................................................................................... 41
Figure 42 Yasuka private commercial and apartment block number 2 is adjacent to block number 1
and was being refurbished at the time of our visit ............................................................................ 41
Figure 43 The Plaza Hotel, which was being refurbished at the time of our visit.............................................. 41
Figure 44 External stairs giving access to the 4th storey of the Plaza Hotel ..................................................... 41
Figure 45 Locations of designated tsunami evacuation buildings in Kesennuma City (Data from
Kesennuma City General Affairs Department). (A) Office building; (B) Yoyoi food factory; (C)
Shoe store; (D) Elderly welfare centre; (E) Tourist Pier and fish market; (F) Prefectural
Office; (G) National government office; (H) Central Community Centre; (I) Kahoku
Newspaper office ............................................................................................................................. 42
Figure 46 An office building in the inner harbour was unused on March 11th. This building sustained
debris damage to the external stairs (front right) which face the harbour. Minor scour
occurred along the seaward face of the building, and around the base of the external stairs. ......... 43
Figure 47 The Yoyoi food factory, where 400 people took refuge. Inundation height was around 8 m
here. ................................................................................................................................................. 43
Figure 48 An elderly persons’ welfare centre which was inundated to around 8 m height. Scour up to 2
m deep occurred on the near side of the building in this photograph. Extensive damage
occurred immediately inland of this building, while 80 survived on the upper floors. ....................... 43

GNS Science Report 2012/17 iv


2012

Figure 49 The large open car park roof space above the fish market provided refuge to 1000 people.
This structure is at the harbour front and inundation almost reached roof level. The open
steel structure enabled tsunami flow at the 1st storey. ..................................................................... 44
Figure 50 The Prefectural Government offices suffered extensive scour and heavy in inundation of 8
m. This building was constructed after 2004 and displays tsunami vertical evacuation
signage at roof level. 200 people survived the tsunami in this building ............................................ 44
Figure 51 The National Government office, where 120 people took refuge from the tsunami. This
building is adjacent to the Prefectural Government offices and was left undamaged when
light steel structures were washed into it. The lower two storeys were inundated. .......................... 44
Figure 52 Kesennuma central community centre, where 450 people took refuge when they evacuated
nearby buildings ............................................................................................................................... 45
Figure 53 The Kahoku Newspaper offices were inundated to at least the ceiling of the 1st storey. 71
people took refuge here. .................................................................................................................. 45
Figure 54 Locations of designated tsunami evacuation buildings in Minami-Sanriku Town: (A)
Matsubara apartments; (B) Takano Kaikan building; (C) Shizugawa Hospital; (D) Fisheries
Co-operative. Significant scour can be seen where seawater is encroaching around building
(A) and to the east of building (D). ................................................................................................... 46
Figure 55 Matsubara apartment block viewed from the seaward side with waterborne debris to the 4th
storey. The low seawall has been destroyed in the tsunami return flow and was sheltered in
part by the apartment building. The foundations of this building are exposed above the water
level due to extensive scour. ............................................................................................................ 46
Figure 56 The Takano Kaikan building 160 m inland of the Matsubara apartment block. Little damage
aside from glazing was observed here, despite heavy damage and collapse of RC buildings
immediately seaward. ...................................................................................................................... 47
Figure 57 Shizugawa Hospital (west building) – the taller of the 2 hospital buildings at 5-storeys, this
section suffered less debris damage than the east building. ............................................................ 47
Figure 58 Shizugawa Hospital (east building) viewed from the seaward side which suffered debris
damage at the 1st and 2nd storeys. ................................................................................................... 47
Figure 59 Minami-Sanriku fisheries co-operative building which was overtopped on March 11th. Image
credit: Google Streetview, August 2011. .......................................................................................... 47
Figure 60 York Benimaru shopping centre, Minato-machi, Ishinomaki City. The rooftop car parking
area is accessible by vehicle ramp and was successfully used for vertical evacuation on
March 11th. ....................................................................................................................................... 48
Figure 61 The Homac hardware store is a single storey building with car parking on the roof, which is
accessible by vehicle ramp. Image credit: Google Streetview, August 2011 ................................... 48
Figure 62 The Hotaru funeral facility. This is an unlikely vertical evacuation facility given the lack of
th
obvious roof space or external access, but successfully provided refuge on March 11 . ................ 49
Figure 63 Kadonowaki Elementary School was damaged by fire while inundated during the tsunami.
People who had sought refuge hear were able to leave the building and evacuate further
uphill at the rear of the school. This building is a typical school structure in Japan; these
buildings are commonly used as evacuation or emergency welfare centres.................................... 49
Figure 64 Locations of designated evacuation buildings in Yuriage, Natori City: (A) Yuriage Junior High
School; (B) Yuriage Elementary School; (C) Yuriage Community Centre. Sendai
International airport is to the south of the area shown here. ............................................................ 50
Figure 65 The Yuriage Community Centre was inundated to the 2nd storey but 43 people survived in
the building. ...................................................................................................................................... 50
Figure 66 Yuriage Elementary School. 870 people took refuge here, where the large roof area and 3rd
rd
storey were above inundation. Direct external access to the 3 storey is available from the
stairs shown here. ............................................................................................................................ 50
Figure 67 Yuriage Junior High School is constructed on ground which is raised 1.8 m above the
surrounding fields. This helped to mitigate damage from the tsunami flow and only contents
damage was observed (EEFIT, 2011). 823 people took refuge here. .............................................. 51
Figure 68 Official vertical evacuation signage, as photographed on the apartment block in Kamaishi. ........... 52
Figure 69 Sign indicating the entrance to external stairways at the apartment block in Kamaishi
(Translation: “Evacuation building entrance (stairways)”.................................................................. 52

GNS Science Report 2012/17 v


2012

TABLES
Table 1 Human perception and indoor/outdoor effects at JMA intensity 5 Lower to 6 Upper. Extract
from Japanese Meteorological Agency (2011b). ................................................................................ 2
Table 2 Correspondence between JMA instrumental intensity and Modified Mercalli intensity (after
Kunugi, 2000). IJMA is JMA instrumental intensity (‘L’ denotes Lower, ‘U’ denotes Upper).
IMM is instrumental Modified Mercalli. IMM1 is Modified Mercalli intensity, which
corresponds to the same values for New Zealand MMI scale of Dowrick (1996). Thresholds
are determined according to the comparison of the description of the intensity at each level. ........... 2
Table 3 Casualty statistics at locations investigated in this report. JMA intensity source: Japanese
Meteorological Agency, 2011c. Total population source: Statistics Bureau, 2010. Casualty
statistics source: Fire and Disaster Management Agency of Japan, 2012 (except Tar Town
– from Ichiro Matsuo, personal communication). These figures do not account for daily
movement of population and fatality rate may have been affected by daily commuting
patterns (after EEFIT, 2011) .............................................................................................................. 3
Table 4 Specifications of tsunami warning/advisory issued by JMA (Source: Japanese
Meteorological Agency, 2006). ........................................................................................................... 8
Table 5 Time sequence of JMA tsunami warnings from March 11th to 13th, indicating the change in
estimated tsunami height for Iwate and Miyagi Prefectures. Between warning #3 and #10,
the major tsunami warning remained in place for these prefectures (Source: Japanese
Meteorological Agency, 2011e). ....................................................................................................... 29

APPENDICES
Appendix 1 Interview questions (Original English version) ................................................................................. 68
Appendix 2 Interview questions (Japanese translation) ...................................................................................... 70
Appendix 3 Summary of key points mentioned in interviews .............................................................................. 74
Appendix 4 Summary of vertical evacuation buildings in locations visited .......................................................... 78

GNS Science Report 2012/17 vi


2012

ABSTRACT

The Great East Japan moment magnitude (Mw) 9.0 earthquake occurred at 14:46 (Japanese
Standard Time) on March 11th 2011. Significant seabed displacement generated the
subsequent tsunami, which caused significant damage in Iwate, Miyagi, Fukushima and
Ibaraki Prefectures. As a result of this event over 19,000 people are dead or missing, with
over 295,000 collapsed buildings along 600 km of affected coastline.

Reconnaissance-level analysis of evacuation preparedness and actions related to the


tsunami has been carried out using semi-structured interviews with local disaster prevention
officials and emergency services officials. Interviews were carried out in Tar Town,
Kamaishi City, funato City (Iwate Prefecture) and Kesennuma City, Minami-Sanriku Town,
Ishinomaki City and Natori City (Miyagi Prefecture). The interviews covered tsunami
awareness, observations and response to natural and informal warnings; style and derivation
of evacuation maps; official warning timing and dynamics; evacuation timing, mechanisms
and issues; and vertical evacuation buildings – availability, designation, public awareness,
utilisation, relationship to maps, and post-event review. The report also presents examples of
hazard and evacuation maps and signs employed in the T hoku region.

Experiences in T hoku during this event are relevant to tsunami mitigation activities in the
State of Washington and in New Zealand, which co-funded this research. These areas have
local earthquake and tsunami risk posed by the Cascadia Subduction Zone and the offshore
Hikurangi subduction margin, respectively. This report provides recommendations for further
development of tsunami mitigation activities in these areas, based on findings from the
interviews.

Overall there was a 96% survival rate of those living in the inundated area of the
municipalities visited. This can be attributed to mostly effective education and evacuation
procedures. Schools education, hazard maps and exercises appear to be the most common
forms of education. Community involvement in planning of evacuation maps, routes and
buildings is common, with many places conducting regular community-level exercises.
Hazard and evacuation maps lacked consistency and both maps and safe locations were
generally designed for a tsunami height that under-represented the worst case scenario.

The natural warning of long ground shaking (reported as more than two minutes, and often
more than three) was widely agreed as enough by itself to have triggered evacuation. Sea
walls reduced effective observation of the natural warning of unusual ocean behaviour in
many places, and fostered a false sense of security in some locations.

Although an early warning system is often seen publicly as key infrastructure in enhancing
tsunami resilience, the expectation of official warnings (and their content) may have slowed
the time taken for people to initiate evacuation in T hoku, compared to if there had been total
reliance on natural warnings. Exposure to previous false ‘major tsunami’ warnings apparently
led to complacency in this event, despite this earthquake feeling much larger than anything
previously experienced. The philosophy of tsunami tendenko was shown to be a positive
education tool which promoted immediate self-evacuation and save many lives.

Peoples’ movements during and after evacuation reveal that many people died unnecessarily
due to delayed evacuation or non-evacuation as a result of social or parental responsibility,
lack of education or scepticism of warnings. Widespread use of motor vehicles caused traffic

GNS Science Report 2012/17 vii


2012

congestion in some areas, when walking, running or cycling would have been much more
effective and saved lives.

Many people returned to the evacuation zone too early in some places because they had not
seen the wave arrive at the expected time given in official warnings, or because they
expected no more waves to arrive. It is critical that people have the awareness that the first
wave may come later than estimated by rapid scientific analysis, and the largest wave may
not be the first.

The evacuation strategy in place at March 11th 2011 was appropriate in that it sent people to
safe locations, used maps and community involvement and was regularly exercised in many
places. Some evacuation centres were not located far enough inland or on high enough
ground because they were not designated using the worst-case tsunami inundation.

There was extensive effective use of both designated and informal vertical evacuation
buildings. The most important considerations for effective use are sufficient height (in relation
to expected inundation depth), reinforced concrete construction, community engagement,
owner agreement, signage, 24-hour access and evacuee welfare. More than one building
owner considered use of their building in evacuation as corporate social responsibility. To
enhance evacuee safety it is prudent to minimise the opportunities for spilled accelerants
such as oil, and debris such as logs in tsunami-prone locations.

KEYWORDS

Great East Japan earthquake and tsunami, evacuation planning, evacuation response,
vertical evacuation, natural warning, tsunami warning.

GNS Science Report 2012/17 viii


2012

1.0 INTRODUCTION

1.1 Background to the event

The moment magnitude (Mw) 9.0 Great East Japan earthquake occurred at 14:46
(Japanese Standard Time) on March 11th 2011. The earthquake mechanism was reverse
faulting at a depth of 24 km on the subduction zone plate interface at the Japan Trench, with
the epicentre at N38.1, E142.9, 130 km offshore of the Oshika Peninsula (Japanese
Meteorological Agency, 2011a). The total rupture area is estimated to be 380 km long and
90-130 km wide (Geospatial Information Authority of Japan (GSI), 2011a). Strong ground
shaking was felt across Japan, and while Japanese Meteorological Agency (JMA) shaking
intensity 7 was recorded at one location (Kurihara City, Miyagi Prefecture), locations in Iwate
and Miyagi Prefectures generally experienced maximum JMA intensities of 5 Lower to 6
Upper (the perception and impacts of which are described in Table 1). This level of shaking
corresponds approximately to Modified Mercalli Intensities 7 to 10 (Table 2), used in New
Zealand and the U.S.

Seafloor geodetic data indicates co-seismic displacement of up to 3 m in an upward direction


(Sato et al., 2011), which was sufficient displacement to generate the tsunami that caused
damage along the entire east coast of Honshu, and particularly significant damage in Iwate,
Miyagi, Fukushima and Ibaraki Prefectures. Coastal towns were afforded a minimum of
around 20-25 minutes between the earthquake and arrival of the first significant (inundating)
tsunami waves. Some areas, such as Ishinomaki City, experienced five significant waves,
the latest of these arriving at least 2 hours and 15 minutes after the earthquake. Many areas
(for example, Natori City and Kesennuma City) remained inundated with standing water until
at least March 12th 2011, with frequent subsequent flooding due to co-seismic subsidence at
the coast of up to 1.2 m (Geospatial Information Authority of Japan (GSI), 2011b). In
Kesennuma, fires that had ignited during the initial tsunami waves continued to burn on
March 12th.

There are some similarities between this event and expectations of a Cascadia tsunami in
Washington State (WA). According to simulations by Venturato et al. (2007), expectations in
Washington are for 1-3 m of seafloor deformation and 1-2 m co-seismic subsidence – similar
to that experienced in T hoku. Tsunami waves are expected to arrive within as little as 20
minutes and the second wave to reach shore has been modelled at 5-6 m height for Ocean
Shores, WA and 7-8 m at Long Beach, WA. These wave heights are sufficient to inundate
the 2nd and 3rd storeys of reinforced concrete (RC) or steel buildings and destroy timber
buildings near the coast.

The extent of destruction that this magnitude of event can cause is highlighted by damage
statistics from the National Police Agency of Japan (2012): 118,810 structures suffered total
collapse, 176,964 partial collapse and 479,429 partial damage in Iwate, Miyagi, Fukushima
and Ibaraki Prefectures. Life loss from this event at February 14th 2012 stood at 19,263
dead and missing (Fire and Disaster Management Agency of Japan, 2012). Fatality rates
were variable between the cities and towns affected – Table 3 shows that the fatality rate for
the locations visited in this research range from 2% to 8%. The highest recorded fatality rate
exceeded 11.5% in Onagawa Town (not visited during this reconnaissance).

GNS Science Report 2012/17 1


2012

Table 1 Human perception and indoor/outdoor effects at JMA intensity 5 Lower to 6 Upper.
Extract from Japanese Meteorological Agency (2011b).

Seismic Human perception and


Indoor situation Outdoor situation
intensity reaction
Hanging objects such as
lamps swing violently.
Dishes in cupboards and
In some cases, windows may
Many people are frightened items on bookshelves
break and fall. People notice
5 Lower and feel the need to hold may fall. Many unstable
electricity poles moving.
onto something stable. ornaments fall.
Roads may sustain damage.
Unsecured furniture may
move, and unstable
furniture may topple over.
Windows may break and fall,
Dishes in cupboards and
unreinforced concrete-block
Many people find it hard to items on bookshelves are
walls may collapse, poorly
move; walking is difficult more likely to fall. TVs
5 Upper installed vending machines
without holding onto may fall from their
may topple over, automobiles
something stable. stands, and unsecured
may stop due to the difficulty
furniture may topple over.
of continued movement.

Many unsecured furniture


It is difficult to remain moves and may topple Wall tiles and windows may
6 Lower
standing. over. Doors may become sustain damage and fall.
wedged shut.

It is impossible to remain Wall tiles and windows are


Most unsecured furniture
standing or move without more likely to break and fall.
6 Upper moves, and is more likely
crawling. People may be Most unreinforced concrete-
to topple over.
thrown through the air. block walls collapse.

Table 2 Correspondence between JMA instrumental intensity and Modified Mercalli intensity
(after Kunugi, 2000). IJMA is JMA instrumental intensity (‘L’ denotes Lower, ‘U’ denotes Upper). IMM
is instrumental Modified Mercalli. IMM1 is Modified Mercalli intensity, which corresponds to the same
values for New Zealand MMI scale of Dowrick (1996). Thresholds are determined according to the
comparison of the description of the intensity at each level.

IJMA 0 1 2 3 4 5L 5U 6L 6U 7

IMM 1 2 3 4 5 6 7 8 9~

IMM1 1 2 3 4 5 6 7 8 9 10 11 ~

GNS Science Report 2012/17 2


2012

Table 3 Casualty statistics at locations investigated in this report. JMA intensity source: Japanese
Meteorological Agency, 2011c. Total population source: Statistics Bureau, 2010. Casualty statistics
source: Fire and Disaster Management Agency of Japan, 2012 (except Tar Town – from Ichiro
Matsuo, personal communication). These figures do not account for daily movement of population
and fatality rate may have been affected by daily commuting patterns (after EEFIT, 2011)

Percentage
Population Dead and Fatality rate
of total pop.
JMA Total residing in missing (at (population
Municipality Residing in
intensity population inundated 14 feb of inundated
inundated
area 2012) area)
area

Tar Town 5L to 5U 4,876 3,000 62% 185 6%

Kamaishi
5U to 6L 39,578 13,164 33% 1,047 8%
City

funato City 6L 40,738 19,073 47% 425 2%

Kesennuma
6L 73,494 40,331 55% 1,368 3%
City
Minami-
Sanriku 6L 17,431 14,389 83% 875 6%
Town
Ishinomaki
6L 162,822 112,276 69% 3,739 3%
City

Natori City 6U 73,140 12,155 17% 966 8%

All areas - 412,079 214,388 52% 8,605 4%

1.2 Goals, methodology and report structure

Many of the issues experienced in terms of evacuation preparedness and response in


T hoku during this event are extremely relevant to tsunami mitigation activities in the State
of Washington owing to local earthquake and tsunami hazard posed by the Cascadia
Subduction Zone, and in New Zealand due to similar earthquake hazard on the offshore
Hikurangi subduction margin.

This research set out to investigate key lessons from Japan that are pertinent to emergency
management and tsunami preparedness in Washington and New Zealand under the
following categories:

• Characteristics of tsunami from a high magnitude local subduction zone earthquake


(Section 1.1)
• Level of tsunami preparedness in affected communities and its role in mitigating life loss
(Sections 2.1 and 3.1)
• Effectiveness of tsunami warnings and evacuation strategies (including maps and signs)
(Sections 2.2 to 2.6 and 3.2 to 3.6)

GNS Science Report 2012/17 3


2012

• Factors affecting emergency response (Sections 2.3 to 2.5 and 3.3 to 3.5), and
• Performance of structures under earthquake shaking and subsequent tsunami loading,
particularly those designed for/designated as vertical evacuation (Sections 2.7 and 3.7)

The research was commissioned by Washington Military Department Emergency


Management Division (Washington EMD) and supported by New Zealand Ministry of
Science and Innovation core geohazards funding to GNS Science, and was conducted in
collaboration with local Japanese colleagues. The report focusses on lessons for
Washington and New Zealand.

This report provides a reconnaissance-level analysis of the themes which arose during
interviews of disaster prevention officials and emergency services officials at seven tsunami-
affected locations in Iwate and Miyagi Prefectures. We discuss these themes in Section 2.0,
and provide recommendations in Section 3.0 for furthering tsunami preparedness and
mitigation in the Washington and in New Zealand. Interview data has been augmented with
discussion of local maps, posters, reports, documents and signs in both Japanese and
English. Appendices 1 and 2 show the interview questions in English and Japanese and
Appendix 3 gives a summary of the key points that were raised during interviews. Appendix
4 provides a summary of vertical evacuation buildings in the locations visited, including
information on construction, signage, inundation on 11th March, and number of people who
took refuge there.

1.3 Interviews

The field survey was carried out from October 19th to 28th 2011 in collaboration with Mr Ichiro
Matsuo of NPO-CeMI, and Professor Hitomi Murakami of Yamaguchi University. Due to the
timing of the survey 7 months after the tsunami, the interviews draw on information gathered
through many surveys carried out by local researchers in the interim period.

Semi-structured interviews with city officials (civil protection, emergency management, fire
department and police department staff) and some local residents were carried out in seven
locations in Iwate and Miyagi Prefectures: Tar Town, Kamaishi City, funato City,
Kesennuma City, Minami-Sanriku Town, Ishinomaki City and Natori City. These locations
comprise variable topography of rias and plains, and among them experienced a wide range
of impacts and evacuation issues on March 11th 2011. A map of the locations investigated is
presented in Figure 1 and a summary of casualty statistics is provided in Table 3.

The focus of interviews was:

• Natural and informal warnings: awareness, observations/experience and response


• Style and derivation of evacuation maps
• Official warning timing and dynamics
• Evacuation timing, mechanisms and issues
• Vertical evacuation buildings – availability, designation, public awareness, utilisation,
relationship to maps, and post-event review

GNS Science Report 2012/17 4


2012

The interview questions were translated into Japanese (see Appendices 1 and 2) and
circulated to interviewees in the days ahead of our meetings. Note that during interviews the
ordering of interview topics naturally fell in the sequence given above, rather than the written
sequence given in Appendices 1 and 2. Sections 2.0 and 3.0 of this report have been
structured according to that natural ordering.

Figure 1 Map of locations investigated in this report, with coastal topography and epicentre of
March 11th main shock indicated. Note the grey lines delineate prefectures. All of the locations
investigated lie within Iwate and Miyagi prefectures (labelled), which received the largest tsunami
wave heights and most of the casualties. Base map credits: GEBCO, NOAA, National Geographic,
DeLorme, Esri.

GNS Science Report 2012/17 5


2012

1.4 Tsunami preparedness in the Tōhoku region of Japan

1.4.1 Previous events and hazard estimation

Tsunami preparedness in Japan is addressed in the Basic Disaster Plan as a subsection of


Earthquake Countermeasures, which is summarised in a plan report from the Cabinet Office
Government of Japan (2007). The Cabinet Office has produced guidelines for the creation of
tsunami hazard maps and designation of tsunami evacuation buildings (Hiroi et al., 2005).
These are followed by prefecture and municipality governments under their responsibility to
formulate and implement local disaster management plans, based on the national plan.

Tsunami preparedness in the T hoku region is borne out of well-recorded impacts of


previous events, primarily the 1896 Meiji Sanriku, 1933 Sh wa Sanriku and the 1960 Chile
tsunami. A technical investigation to define characteristics of expected earthquakes of the
Japan Trench was carried out between 2003 and 2006, in which the expected damage from
these events was estimated and necessary mitigation strategies developed. Eight
earthquake scenarios were examined for the northeast of Japan, and their estimated
impacts used to develop strategies for mitigation of economic damage and life loss (Central
Disaster Prevention Council, 2008). Analysis of these eight separate events formed the
basis of hazard assessment from Japan Trench tsunami until early 2011, with no
consideration for the rupture of multiple segments at once (Expert Committee on the
Working Group on Tsunami disaster evacuation, 2010), as occurred on March 11th.

Preparedness prior to March 2011 focussed on the locally recurrent Miyagi-ken-oki


(“Offshore Miyagi Prefecture”) tsunami, which has been consistently used as the basis for
numerical tsunami simulation, hazard mapping and determination of the design-level
tsunami for structural defences. The Miyagi-ken-oki earthquake is a recurrent earthquake
event of around magnitude 7.5 with a source area off the Oshika Peninsula, Miyagi
Prefecture (in the same area as the epicentre of March 11th event, Figure 1). This size of
earthquake has occurred five times since 1835, the most recent occurrences in 1978 and
2005, with no significant tsunami. The event was estimated to have a mean recurrence
interval of 35.2 years and an estimated 97% probability of occurrence in the period 2002-
2032 (Ohtake and Ueda, 2002). Due to the frequency of this type of event and knowledge of
the source area, this scenario is the one most often simulated to map tsunami hazard, while
inundation heights and extents of the Meiji and Sh wa era tsunami are often also delineated
on hazard maps (see Section 2.2).

The absence of the 869 Jogan Tsunami in informing preparedness activities is surprising
given that it has been demonstrated inundation extended more than 4 km inland (Minoura,
Imamura, Sugawara, Kono, and Iwashita, 2002). This extent is similar to the inundation
extent on March 11th 2011 in the same area. This extreme event was not included in hazard
modelling and mapping due to it being perceived as an outlier (Earthquake Engineering
Research Institute, 2011a), and perhaps also due to a lack of knowledge of the potential
source area and absence of tsunami deposit data on the inundation extent on other sections
of the T hoku coast.

1.4.2 Hazard estimation and mapping

Hazard estimation is consistent at prefecture level as it is the prefectural government who


carry out the tsunami estimation for all coastal municipalities within their prefecture. The

GNS Science Report 2012/17 6


2012

delineated hazard is provided to the individual municipality governments, who produce


hazard maps and evacuation maps as part of their local disaster management plan.
Community engagement with evacuation mapping is discussed in Section 2.1, which
includes examples of hazard maps from the locations visited. The inconsistencies between
municipalities in representation of hazard and evacuation routes or refuges are shown in
Section 2.2 and in the Miyagi Prefecture online database of maps (Miyagi Prefectural
Government, 2004).

1.4.3 Education and evacuation exercises

Regular education and evacuation drills are conducted by municipal governments to


maintain disaster prevention awareness. A national “Disaster Reduction Day” is held
annually, during which every prefecture carries out large-scale decision-making simulations.
There is also an annual tsunami evacuation drill held in Miyagi and Iwate prefectures on the
anniversary of the 1960 Chile tsunami. In addition to this, disaster prevention activities are
carried out by local groups at the city or neighbourhood level for tsunami in addition to other
perils.

The philosophy of tsunami tendenko relies on the trust that every person in a community
knows what to do in the event of a tsunami to self-evacuate immediately (rather than helping
or waiting for others) and get to an evacuation refuge according to a home or work
evacuation plan. This philosophy is described very well in a video documentary by D.
Harding and B. Harding (2011).

Despite these exercises and the well-known history of tsunami in Japan, a review of
evacuation in Kesennuma City following a local offshore earthquake in 2003 (causing JMA
shaking intensity up to 6 Lower) showed that almost 86% of residents, including those in the
tsunami hazard zone, did not evacuate (Suganuma, 2006) and a low rate of evacuation also
occurred following a tsunami warning in 2004 in the Tokaido region. Suganuma (2006)
describes a decline in disaster prevention awareness over time since the 1995 Hanshin-
Awaji (Kobe) earthquake. A survey of 1,808 residents in areas of Sendai subject to a
tsunami warning on May 28th 2010 (following the Maule, Chile earthquake) shows only
53.9% of residents evacuated (City of Sendai, 2010).

1.4.4 Earthquake early warning system and rapid official tsunami warnings

Another high-profile tool in the preparedness framework for Japan is the JMA-operated
earthquake early warning system. A network of seismometers allows rapid detection of
earthquakes and analysis of the magnitude and hypo-central location. P-wave arrival time
near the epicentre allows estimation of arrival for the more damaging S-waves. This system
can provide tens of seconds warning ahead of damaging ground shaking, allowing time for
people to find cover, and also facilitates rapid estimation of tsunami generation.

Within 2-3 minutes of a tsunamigenic earthquake occurring, estimated tsunami wave heights
and arrival times can be provided in a tsunami advisory or warning, based on the estimated
earthquake magnitude (Japanese Meteorological Agency, 2011d; Ozaki, 2011). Tsunami
warnings are issued to 66 individual regions by JMA with a level of severity determined by
expected tsunami height at the coast (Table 4). Observations at offshore GPS buoys are
used to revise earlier estimates as water level or pressure changes at the buoy (Ozaki,
2011). JMA disseminates advisories and warnings to TV and media, and to local
governments using a combination of dedicated online, radio, and satellite wireless networks.

GNS Science Report 2012/17 7


2012

Municipality governments are then responsible for issuing evacuation orders to the public on
receipt of a warning from JMA or national government (Suganuma, 2006), which is achieved
via roof-top or street-based loud speakers and indoor receivers.

Table 4 Specifications of tsunami warning/advisory issued by JMA (Source: Japanese


Meteorological Agency, 2006).

Category Indication Forecast tsunami height


Forecast heights are specifically
Major Tsunami height is expected to be 3 indicated for every region; namely
Tsunami tsunami meters or more. 3 m, 4 m, 6 m, 8 m and 10 m or
Warning more.
Tsunami height is expected to be
Tsunami Same as above, but 1 m or 2 m.
up to 2 meters.
Tsunami height is expected to be
Tsunami Advisory 0.5 m
about 0.5 meters.

1.4.5 Coastal defences

‘Hard’ coastal defences are a common feature of the Japanese coastline, and at a given
location may include a combination of offshore breakwaters, groynes, seawalls (at the
shoreline, or further inland), concrete revetments and river-mouth flood gates. The design-
level tsunami for these defences is often those that have been previously experienced:
Kamaishi breakwaters are based on the 1896 Meiji tsunami; the Tar seawall was
constructed as a result of the effects from the 1896 and 1933 tsunami. Other defences are
designed according to the Miyagi-ken-oki tsunami scenario.

1.5 Tsunami preparedness in the State of Washington and New Zealand

The authors have previously undertaken numerous social research projects on tsunami
preparedness and evacuation in United States (specifically the State of Washington), New
Zealand and Japan. These projects have been in collaboration with local and national level
emergency managers in New Zealand, Washington EMD, and CeMI (an environmental
planning non-profit organisation) in Japan. They have included studies of public perception,
tourism preparedness, warning system effectiveness, evaluation of evacuation plans, and
standardisation of mapping and signage.

1.5.1 Washington

For over a decade the State of Washington has developed a programme of tsunami
understanding and preparedness, through a combination of state initiatives and wider
national programmes. This occurred in response to the increased awareness among
scientists of the tsunami risk posed by the earthquakes along the Cascadia Subduction
Zone, offshore of the Pacific Northwest states. As part of an integrated evaluation program,
Johnston et al. (2005) carried out several surveys and focus groups between 2001 and 2003
to evaluate tsunami preparedness in Washington. This work found that earlier initiatives had
been “moderately to highly effective in raising public awareness of the tsunami hazard”, but
that this awareness did not necessarily translate into preparedness, which was recorded as
“low to moderate”. Awareness among visitors was much lower than that of residents, and
Johnston et al. (2007) followed up these findings with an assessment of preparedness in the
tourism sector. This work, carried out in 2005, found a low level of staff training on tsunami
hazards, and little provision of hazard information to guests. In response to the findings of

GNS Science Report 2012/17 8


2012

Johnston et al. (2005), various outreach initiatives and educational tools were developed to
increase preparedness in Washington, while staff training, workshops and a guidebook were
developed specifically for the tourist sector in response to the findings of Johnston et al.
(2007). The most recent follow-up research in this series of evaluations, Johnston et al.
(2009), noted that high staff turnover remains a challenge to achieving high levels of
preparedness in the tourism sector although there have been improvements. It was also
recognised by Johnston et al. (2009) that improvements have been made in the provision of
loudspeaker warning systems and community-based evacuation drills, although participants
from both Ocean Shores and Long Beach expressed concern over the available evacuation
routes. Further evaluations of siren tests and community-based evacuation exercises were
carried out in September 2010; the findings showed mixed levels of siren audibility in the
current system, but highlighted the effectiveness of siren and evacuation drills as a means of
hazard education (Leonard et al., 2011).

The Washington EMD program focusses on education and preparedness of both the
resident public and the tourism sector. This is achieved in part through public forums, town
hall meetings in collaboration with local Emergency Managers, and training of local
volunteers to serve as community tsunami educators. Tsunami information, preparedness
and evacuation educational materials are provided via internet, brochures, fact sheets,
video, resource guides, guidebooks, childrens’ cartoon books, trivia sheets, games, and
evacuation maps are available (http://www.emd.wa.gov/hazards/haz_tsunami.shtml).
Evacuation signage is widely deployed in tsunami hazard zones.

Tsunami warning infrastructure includes All Hazard Alert Broadcast (AHAB) to radios and
free-standing siren towers and NOAA weather radios are provided to low income families
and schools. Twelve communities in Washington are recognised as ‘Tsunami Ready’ under
the National Oceanic and Atmospheric Administration/National Weather Service
(NOAA/NWS) ‘Tsunami Ready’ program, which helps communities increase tsunami
preparedness (NOAA/NWS, 2011).

The provision of evacuation maps is a major initiative covering over 20 different locations.
Some maps use three zones of ‘greatest risk’, ‘marginal risk’ and ‘higher ground’, evacuation
routes, assembly areas, roads, rivers and place names (Figure 2). Other maps use only two
zones: ‘tsunami hazard zone’ and ‘outside of hazard zones’, for example in Clallam County
(Figure 3). These maps are provided with additional information as part of a short brochure.
In the planning of evacuation routes and assembly areas, Washington EMD and the
Department of Natural Resources assess landslide and liquefaction hazards which may
impede evacuation.

There are inconsistencies between maps produced by different state authorities in the U.S.
and Kurowski, Hedley, and Clague (2011) discuss this issue in their evaluation of tsunami
evacuation maps in Washington. New National Tsunami Hazard Mitigation Program
(NTHMP) guidelines have been released to ensure there is enhanced consistency in this
respect from January 2012 (NTHMP Mitigation and Education (MES) Subcommittee, 2011).

GNS Science Report 2012/17 9


2012

Figure 2 Excerpt from the tsunami evacuation map Clallam Bay, Clallam County, Washington.
This map uses 2 zones, local sites of importance (fire department, etc.), evacuation routes and
assembly areas (Clallam County Emergency Management, n d).

Figure 3 An example evacuation map in


Washington. This map is for Long Beach and
Ilwaco, Pacific County, and illustrates the use
of 3 zones, evacuation routes and assembly
areas (Pacific County Emergency
Management, 2007).

GNS Science Report 2012/17 10


2012

Signs for hazard zones and evacuation routes are standardised along the Pacific Coast, with
signs indicating the tsunami hazard zone with brief instructions regarding natural warnings
(Figure 4) and evacuation route signage (Figure 5).

Figure 4 ‘Tsunami Hazard Zone’ sign with instructions to evacuate to high ground or inland in
response to an earthquake.
Figure 5 ‘Tsunami Evacuation Route’ sign with arrow depicting the direction of evacuation. These
signs are placed regularly along roads in tsunami hazard zones.

1.5.2 New Zealand

The risk of tsunami from the offshore plate boundary has been apparent to scientists in New
Zealand for well over a decade, but political and public traction for mitigation only increased
following the Indian Ocean tsunami in 2004. Guidance for New Zealand is now in place for
tsunami warning (MCDEM, 2010a), evacuation mapping (MCDEM, 2008a), signage
(MCDEM, 2008b), land use planning (Saunders et al., 2011), and guidance for official CDEM
messages (MCDEM, 2010b).

The standard base map for evacuation maps shows three evacuation zones: a coastal
exclusion zone that can be placed off limits when only the foreshore is at risk, an orange
zone that intends to cover most distant sources, and a larger yellow zone that includes the
local and regional sources (Figure 6). The maps are then developed in discussion with the
local community with respect to what is depicted on the map – e.g. road names and key
infrastructure.

Standard signs include an ‘evacuation zone sign’, information boards which include the
evacuation map, various ‘evacuation route’ signs, an ‘evacuation safe location’ sign, and a
‘previous event’ sign (which is unused and may be re-considered in light of T hoku events).
The first two signs are shown on the information board (Figure 7), while the latter two have
not yet been implemented in New Zealand.

Neither Washington nor New Zealand employs large amounts of ‘hard’ coastal defences for
coastal protection – sea defences are typically placed in ports and are designed to protect
against storm waves, not tsunami waves. Beach and dune environments dominate the
shoreline in low-lying areas of both regions.

GNS Science Report 2012/17 11


2012

Currently there are no early warning systems in place for Washington or New Zealand for
local source tsunami. Local tsunami warning would be natural and informal warnings,
whereas regional to distant tsunami allow warnings to be disseminated through official
channels. It is known that there is public misunderstanding that official warnings and warning
hardware (e.g. sirens) will cover local sources – this is an ongoing issue for public
awareness and preparedness.

Figure 6 An example tsunami hazard map from Whananaki, New Zealand.


Figure 7 An example tsunami evacuation map and information board as displayed outdoors (1.2
m wide) in the Island Bay community, Wellington, New Zealand.

GNS Science Report 2012/17 12


2012

2.0 INTERVIEW OUTCOMES

This section discusses emergent themes from the interviews with disaster prevention and
emergency service officials in the T hoku region, and relates these issues to Washington
and New Zealand. Each sub-section contains information which has some cross-over to
other themes, and several themes may be inter-related or influenced by the same factors.
No theme or event highlighted here should be considered in isolation. A summary of the key
points raised in interviews at each location is presented in Appendix 3 for quick reference.
Recommendations for Washington and New Zealand are given separately in Section 3.0 in
consistent sub-sections.

2.1 Community involvement in evacuation planning and preparedness

Community disaster groups are common in Japan and many cities and towns have multiple
disaster prevention groups who are engaged with local government in awareness and
education activities at the neighbourhood level.

High levels of community engagement were apparent in most of the interviews, particularly
with residents being involved in production of evacuation maps in Kamaishi, funato,
Kesennuma, and Minami-Sanriku. In these locations, the prefectural government provides
hazard mapping to the municipal government who then work with community groups to
develop the evacuation maps and identify suitable (or historical) places of refuge. This
approach is not ubiquitous – in Natori City, the development of evacuation mapping was said
to be the responsibility of local government who then provide the finalised evacuation map to
the public.

The local community was involved in identification of potential buildings for vertical tsunami
evacuation in funato, Kesennuma and Kamaishi. Community members had either
approached building owners about the potential use of a building, or approached municipal
government with suggestions of potential buildings. There have been mixed outcomes
regarding the eventual official designation of buildings where the community have been
involved in this process, but it is clear that community engagement in vertical evacuation
strategies is beneficial, as discussed in Section 2.7.

In addition to the nationwide tsunami disaster drill carried out annually on the anniversary of
the 1960 Chile tsunami, community groups in Natori City, Ishinomaki City and Miyako City
carry out more regular drills for tsunami, as well as other hazards. The officials we
interviewed were not able to provide data on the levels of resident participation in such drills.

Community involvement in planning and preparedness is well-established in Washington,


where the Emergency Management Division engages with coastal communities through
county level emergency management groups, who co-ordinate community education
sessions and evacuation drills. There has previously been a high level of community interest
in participating in focus groups on tsunami risk and preparedness, with proactive preparation
among many hotels in coastal Washington (Johnston et al., 2005). An AHAB siren test and
evacuation drill on September 21st 2011 achieved 100% participation of schools in Grays
Harbor County, although participation was lower in other counties (J. D. Schelling, personal
communication, December 17th 2011), illustrating the high levels of engagement which have
been achieved. Project Safe Haven is a current vertical evacuation strategy development

GNS Science Report 2012/17 13


2012

project, in which the community drives discussion of conceptual facility locations, design of
facilities, and development of a preferred strategy (Project Safe Haven, 2011a, 2011b). This
project provides raised awareness in the coastal communities and is an excellent model for
developing and enhancing community-based evacuation strategies.

In New Zealand, national guidelines and standards for maps and signs have existed since
2008, but the application of these guidelines at community level is just developing
momentum in 2012. Locally specific evacuation maps have been developed across several
New Zealand regions, with varying approaches to community engagement. Most local
authorities at least consult the community on appropriate evacuation routes at community
meetings. Whangarei District and Wellington City have developed maps with some
communities as part of all hazard community response planning, with full ownership of
maps, plans, sign locations and exercise planning lying with the community planning group.
Evacuation drills presently only occur in a few communities nation-wide. Civil Defence and
Emergency Management (CDEM) Groups of emergency services and disaster managers
are mandated through legislation in all regions, and community-based planning for tsunami
is expected to continue to grow in coming years.

2.2 Hazard/evacuation maps and signs

2.2.1 Hazard/evacuation maps

Maps depicting tsunami hazard (with or without evacuation facilities and routes) had been
developed prior to 2011 for all of the communities visited during this research. The content
and style of evacuation maps therefore varies greatly between municipalities in terms of both
hazard used and features pertaining to evacuation. Community involvement in defining
evacuation routes was a feature of map development in Kamaishi, funato, Kesennuma and
Ishinomaki, but in Natori City the definition of refuges and evacuation routes was carried out
by the municipal government.

Although a full evaluation of evacuation maps is outside of the scope of this report, notable
variations in the maps for locations visited (Figure 8 to Figure 14, obtained from local officials
or government websites) are summarised below.

The inclusion of previous or modelled events is inconsistent. To delineate the tsunami


hazard, most municipalities map recorded inundation heights or extents of one or more of
four key tsunami scenarios: the 1896 Meiji tsunami, 1933 Sh wa tsunami and 1960 Chile
tsunami, and the modelled extent or heights of tsunami from the expected Miyagi-ken-oki
earthquake. The hazard map for Ishinomaki City used the modelled Miyagi-ken-oki
earthquake and the official recognises that the consideration of source events in their
approach was not robust enough. It appeared that several emergency managers interviewed
may not have had a clear idea of which data originated from modelling, and which
represented historic events, but this may have been uncertainty related to language
translation. We were unable to discuss clarity of the maps with members of the public.

Hazard or evacuation zone style varies between maps. Some maps use variable colouring to
indicate modelled inundation depth (Figure 10, Figure 11, Figure 13); others provide only a
single colour delineation of inundation extent in historic events (Figure 8, Figure 12).

GNS Science Report 2012/17 14


2012

Generally maps appear quite crowded with a range of topographic, cultural as well as
tsunami evacuation related content. In some cases additional historic event lines are present
as well as coloured zones. However, the location of at least one tsunami hazard zone stands
out well in all maps.

All of the evacuation maps that we observed show refuge locations although, the symbols
used are inconsistent between maps. Definitions of refuges vary and include generic ‘safe
refuge’, ‘tsunami-specific refuge’, ‘vertical evacuation building’ (Figure 11) or ‘welfare
location’s, with some maps using different symbols to define refuge type. Many of the
refuges shown on maps were designated based on community function and capacity and
were typically schools or community centres. During the evacuation on March 11th many
people evacuated directly to designated shelters.

There is limited illustration of evacuation routes on the maps. funato (Figure 10) and
Kesennuma (Figure 11) use arrows indicating the approximate best route out of the
inundation zone, but there is no prescriptive route marked. Other maps do not include any
graphical indication of evacuation route.

Text-based evacuation or preparedness information is provided on most maps, but the


content of this information varies. The Tar Town map includes information on previous
tsunami events, while the maps for funato and Natori have advice on preparedness and
space for residents to fill out important information.

The interviews carried out in this study did not provide sufficient insight into the impact of
inconsistent evacuation maps on evacuation on March 11th 2011. This requires further
research as Japan amends and enhances its hazard modelling and evacuation mapping in
light of the T hoku tsunami. Recent work by Kurowski et al. (2011) highlighted that some
similar inconsistencies occur in evacuation mapping in Washington and Oregon (see Section
1.5), and any research in Japan into the impacts of variable evacuation maps on the
response of transient populations or different levels of risk perception in Japanese coastal
communities could provide additional guidance as to enhancing the approach taken in
Washington. Consistency in evacuation mapping is already recognised as an issue in the
United States and all maps produced after January 1st 2012 are subject to guidelines set out
by the NTHMP Mitigation and Education Subcommittee (2011) to ensure minimum
requirements in evacuation mapping. Updating of current maps to ensure consistency is
expected to be complete by the end of 2012 (J. D. Schelling, personal communication,
March 9th 2012).

Consistent evacuation mapping in New Zealand is being achieved by following the


guidelines set out in MCDEM (2008a). Some variation still exists in terms of the delineation
of the location of evacuation zone boundaries, where local CDEM group members are
responsible for producing their own mapping of zones from hazardous wave heights e.g.
Hawke’s Bay CDEM Group (2011). In New Zealand there is a strong emphasis on simple
maps showing zones, routes and safe locations, and as concise as possible messaging
focussing on the required response to natural and informal warnings (evacuate all zones) vs.
official warnings (the warning will state the zone(s) to evacuate). One community group, in
conjunction with Wellington Emergency Management Office, has delineated the beginning of
the mapped safe zone with a blue line painted across roads throughout the suburb. This has
generated very high public awareness and is planned to be rolled out city-wide.

GNS Science Report 2012/17 15


2012

Figure 8 Extract from Tar tsunami hazard map, indicating three sections of tsunami wall (red
(1958 section), black and yellow thick lines), tsunami gates (blue triangles and circles). The inundation
extent of the 1986 Meiji (thin red line) and 1933 Sh wa tsunami (thin blue line). Evacuation refuges
are marked as red triangles and welfare centres as red circles. Source: Tar Town Disaster
Prevention Office.

Figure 9 Kamaishi City evacuation map (Source: Kamaishi City Disaster Prevention Office).

GNS Science Report 2012/17 16


2012

Figure 10 funato City evacuation map (Source: funato City Disaster Management).

GNS Science Report 2012/17 17


2012

Figure 11 Excerpt from Kesennuma City evacuation map (Source: Kesennuma General Affairs
Department).

GNS Science Report 2012/17 18


2012

Figure 12 Minami-Sanriku hazard map with evacuation refuge locations indicated by numbered red
circles (Miyagi Prefectural Government, 2004).

Figure 13 Excerpt from the Ishinomaki tsunami hazard map with evacuation refuges (Ishinomaki
City, n d).

GNS Science Report 2012/17 19


2012

Figure 14 Natori City hazard map (Source: Natori City Disaster Prevention Office).

2.2.2 Hazard/evacuation signs

Road signs indicating evacuation routes or extent of inundation zones are inconsistent
between prefecture and municipality, however, they form a key part of tsunami preparedness
and are commonly seen. This sub-section demonstrates the range of information displayed
on signs in Iwate and Miyagi.

Road signs indicating the estimated extent of tsunami inundation zones (Figure 15 and
Figure 16) are present along the coast in Kamaishi district. On roads out of Tar , similar
signs mark the extent of inundation in the 1933 Sh wa tsunami.

Figure 15 A sign displayed above the road indicating to drivers that they are entering an estimated
tsunami inundation area. These signs were seen on roads within Kamaishi district.
Figure 16 A sign displayed above the road indicating to drivers that they are exiting an estimated
tsunami inundation area. These signs were seen on roads within Kamaishi district.

GNS Science Report 2012/17 20


2012

Evacuation route signs in Miyako City and Matsushima Town commonly include the distance
to safe places, in addition to a directional arrow. There is consistent use of the green symbol
depicting evacuation to high ground (Figure 17 to Figure 20). Matsushima Town also has
information boards displaying a schematic map of the entire evacuation route and location of
the designated assembly point on the evacuation route (Figure 19) in addition to signs
painted onto the pavement in town (Figure 20). This style is not consistent for all signs in
Matsushima, however: Figure 21 shows a more basic sign indicating just the direction of
evacuation. The inclusion of distance on evacuation route signs is not currently included in
signage standards for New Zealand (MCDEM, 2008b) or Washington.

Figure 17 Tsunami evacuation route direction and distance sign displayed on a sign in
Hitachihamacho, Miyako City.
Figure 18 Tsunami evacuation route direction and distance sign displayed on a lamp-post in
Hitachihamacho, Miyako City.

Figure 19 Tsunami evacuation route information board and directions (with distance to safe
location) in Matsushima Town, Miyagi Prefecture.
Figure 20 Tsunami evacuation route sign with direction and distance to safe location painted on the
pavement in Matsushima Town, Miyagi Prefecture.

GNS Science Report 2012/17 21


2012

Figure 21 A sign in Matsushima Town, displaying the tsunami evacuation route.


Figure 22 Tsunami evacuation route sign with solar powered lamp, on a road leading to high
ground out of Tar Town.

Tar Town has signs posted regularly at short distances (20-30 m) along the road out of the
Town towards high ground. These signs are mounted on posts with solar powered lamps for
night-time use (Figure 22).

funato City uses signs indicating the walking route direction and distance to safe refuges,
providing the name of the refuge in English and Japanese (Figure 23). The city also uses
signs which light up at night.

Figure 23 Evacuation route sign in funato City, indicating the distance to a “tsunami emergency
place of refuge”. Photo courtesy of EEFIT.

In addition to the information boards and pavement-mounted signs in Matsushima Town, we


also observed an example of natural warning education in the centre of town, situated close
to several tourist sites (Figure 24). This sign displays the standard green icon representing
evacuation to high ground.

Another style of tsunami warning sign was observed on a road out of Kamaishi (Figure 25);
this symbol is consistent with other signs observed in the district in terms of the yellow and

GNS Science Report 2012/17 22


2012

black icon used, but this symbol not seen in any other districts, which appear to only use the
green evacuation symbols. Both the green and yellow symbols have been accepted as ISO
international standards after being proposed by Japan. However, both New Zealand and
Washington use a blue sign style very similar to each other, which is common in other
countries. All signs discussed above have bilingual text for at least the basic information,
enhancing understanding further than recognition of the tsunami symbol for people with little
understanding of Japanese text.

A muster point for use in evacuation of children was seen in Hitachihamacho, Miyako City
(Figure 26, Figure 27). This sign is located on the wall of a house on the road at the base of
a hill. The owner of the house told us that children are instructed to gather at this sign and
evacuate up the hill together.

Figure 24 A sign in the centre of Matsushima Town, advising evacuation on experiencing an


earthquake.
Figure 25 A sign indicating tsunami hazard, on a road out of Kamaishi City. Translation reads:
“Tsunami Attention! Tsunami inundation hazard zone”.

Figure 26 Tsunami evacuation muster point for children in Hitachihamacho, Miyako City.
Figure 27 Close-up view of tsunami evacuation muster point for children in Hitachihamacho,
Miyako City. Translation reads: “Tsunami shelter (Kuwagasaki elementary school)”.

Signs or plaques marking inundation depth in previous tsunami were seen in Minami-Sanriku
(Figure 28 and Figure 29) and funato City (Figure 30). These markers do not indicate
maximum inundation extent in the 1960 Chile event; rather indicate depth at a particular
location. The marker in Figure 28 indicates the tsunami inundation depth was 2.4 m at this

GNS Science Report 2012/17 23


2012

location in the 1960 Chile tsunami; inundation depth at this point was at least 10 m on March
11th 2011. It was not clear from interviews with emergency officials, whether these signs
caused under-estimation of tsunami hazard in the local population, but the authors believe
that this sort of demonstrative information should be reserved for maximum events, and in
the case of these locations should have more appropriately reflected the larger 1896 or 1933
tsunami flows if known. Ancient stones have been reported in the media, which indicate the
extent of historic tsunami; these were not observed by the authors.

Figure 28 A marker post indicating past inundation height in central Minami-Sanriku Town. The
Japanese text gives notification that the Shizugawa hospital is a designated vertical evacuation
building. The Japanese text informs people to evacuate to designated evacuation places such as
parks, highlands, and Shizugawa elementary school and Shizugawa hospital.
Figure 29 A marker post indicating past inundation height in central Minami-Sanriku Town. This
marker was found close to a care home (1 km inland), and displays the standard icon representing
evacuation to high ground.

Figure 30 Tsunami height marker on a building wall in funato City, indicating the maximum
inundation height at this point during the Chile tsunami on May 24th 1960. Photograph courtesy of
EEFIT.

GNS Science Report 2012/17 24


2012

In Ishinomaki, we observed a map which focussed on the local landslide hazard and
evacuation routes designed for landslide only (Figure 31). This map shows evacuation
routes to shelters which are away from steep hillsides, but encourage evacuation into areas
close to the banks of the Kitakami River. On March 11th 2011, the specified evacuation
shelters were inundated by tsunami flow from the river channel. This illustrates the need for
information boards and signs to include multiple hazards where this is appropriate.

Figure 31 Landslide hazard and evacuation map on an information board in Ishinomaki City.

2.3 Response to natural and informal tsunami warnings

2.3.1 Natural warnings

The officials interviewed in Tar , Kesennuma, Minami-Sanriku and Natori reported that
ground shaking lasted longer than 2 minutes (most reported as long as 3 minutes) and in
Kesennuma, Minami-Sanriku and Ishinomaki the interviewees acknowledged that this
earthquake felt unusual or stronger than they had ever experienced previously (including
previous Miyagi-ken-oki events in 1978 and 2004). Most interviewees, when questioned
further, felt that they could have stood up in the earthquake and that ‘stronger’ generally
referred to the length of shaking and the long and variable period of the shaking. As a result
of the ground shaking, they recognised that this could be the recurrent Miyagi-ken-oki event
and there was recognition of the potential for tsunami. The JMA shaking intensity at each
location is given in Table 3. Drawdown of the tsunami was only cited in our interview at
Minami-Sanriku.

Public response to the natural warning of strong and long ground shaking was variable. In
Tar , most people in the town are believed to have evacuated in response to ground shaking
and although an audible tsunami warning was broadcast by the disaster prevention officials,
this was broadcast while people were already running. Tar Town officials describe a history
of strong communication of tsunami risk from parents to children, with a tradition of telling
stories about tsunami, education in school on tsunami including communication on the
subject between local and visiting schoolchildren. The official himself told us how he was
often told as a child to “run straight away, even if the ground shaking is small” in an
earthquake. Local tradition in Minami-Sanriku encourages evacuation from coastal areas

GNS Science Report 2012/17 25


2012

when an earthquake occurs or when drawdown of the sea occurs. It appears that in both
Tar and Minami-Sanriku traditions and education were heeded as they suffered fatality
rates of only 6% despite inundation of more than 10 m.

A study of 113 residents in Kamaishi showed that 66.4% of the population evacuated
immediately, and 72% of those evacuated directly as a result of the natural warning (NPO
CeMI, 2011). This immediate action is a result of education and previous experience.
Tsunami education efforts by local government stress the importance of natural warnings
and this event has re-affirmed to the disaster prevention office in Kamaishi the importance of
this education.

Local disaster prevention officials believe that “most people” evacuated coastal areas of
funato as a result of ground shaking, while it is estimated that 20-30% did the same in
Kesennuma (the majority, however, waited for an official “formal” warning before
evacuating). The response of many people to natural warnings – overwhelmingly in this case
being the strength and especially duration of ground shaking – can be attributed to the
history of significant tsunami impacting the rias locations, and the resultant education to run
to high ground in the case of strong ground shaking.

Unfortunately, the immediate evacuation of low-lying areas was not ubiquitous. Ground
shaking did trigger evacuation in Ishinomaki, although this appears to be in response to
education for earthquakes rather than tsunami, and people evacuated to local parks, rather
than specifically uphill. Evacuation training and schools disaster education in the city was
focussed on earthquake events without consideration for tsunami.

In Natori, young people were said to have evacuated inland on feeling the strong ground
shaking (JMA intensity 6 Upper) but many older people did not believe a tsunami would
affect them from this earthquake, as they had no experience of tsunami reaching the
residential areas after previous earthquakes. The survey by NPO-CeMI shows that 57.1% of
evacuees interviewed did evacuate immediately, although there was only 56.7% of those
who evacuated due directly to the natural warnings (NPO CeMI, 2011). Due to the lack of
recognition of subsequent tsunami hazard, 30% of residents stayed to clear up earthquake
debris, and 40% tried to ensure the safety of family and neighbours. As result, 54.5% of
those who delayed evacuation evacuated between 20-60 minutes after the earthquake.

The immediate reaction of large numbers of people to ground shaking has a cumulative
“cascading effect” on increasing the overall evacuation rate, as other people (who may be
confused or unsure of what action to take) see people running and as a result begin to
evacuate. This effect was reported in Kamaishi, Natori and funato.

Current advice in New Zealand for local tsunami, which are not expected to provide sufficient
time to an issue official warning, is that:

“Persons in coastal areas who experience strong earthquakes (hard to stand up); experience
weak earthquakes lasting for a minute or more; or observe strange sea behaviour such as
the sea level suddenly rising and falling, or hear the sea making loud and unusual noises or
roaring like a jet engine should not wait for an official warning. Instead, let the natural signs
be the warning. They must take immediate action to evacuate predetermined evacuation
zones, or in the absence of predetermined evacuation zones, go to high ground or go
inland.” (MCDEM, 2010a)

GNS Science Report 2012/17 26


2012

This advice is appropriate for the reported experiences in Iwate and Miyagi on March 11th
2011 and it is encouraging that for a similar event in New Zealand, the current education is
appropriate.

Likewise, advice given by Washington EMD and West Coast and Alaska Tsunami Warning
Centre (WCATWC) instructs people to evacuate coastal areas immediately when ground
shaking is felt, although there is no clear qualification (stating what sensation people might
experience) of the level of ground shaking that might be followed by tsunami:

“If you feel the ground shake, evacuate inland or to high ground immediately and return only
after officials say it is safe to do so.” (Washington Military Department Emergency
Management Division, 1999)

“A strong earthquake felt in a low-lying coastal area is a natural warning of possible,


immediate danger. Keep calm and quickly move to higher ground away from the coast.”
(West Coast and Alaska Tsunami Warning Center, 2011)

This is in contrast to the unsuitable advice given on the FEMA ready.gov website:

“If an earthquake occurs and you are in a coastal area, turn on your radio to learn if there is
a tsunami warning.” (FEMA, 2011)

2.3.2 Informal warnings

In addition to peoples’ reaction to seeing others evacuating, informal warnings in other forms
were effective in saving lives. These warnings include hotel owners instructing or guiding
guests to high ground (e.g. in Tar and Unosumai, Kamaishi). In funato, the disaster
prevention official cited business owners who warned employees to evacuate and a resident
who drove north up the valley to warn his colleagues. At Minato High School in Ishinomaki, a
teacher used the school announcement system to warn the area around the school that a
tsunami of over 6 m was coming; it is estimated that over 100 people were saved due to this
warning.

Events in Natori City show that informal warnings (or confirmation of warnings) from the
media cannot be relied on following a strong local earthquake. Tsunami arrival at Natori was
60 minutes after the earthquake, and around 30 minutes after arrival of the first waves at
locations further north. Due to power outages, television and telephones were not
functioning, so residents could not receive information that these locations that had been
affected, thus indicating the possibility of Natori being affected. Residents did talk to one
another and encourage others to evacuate, but many were still reluctant to leave.

Given the potential for electronic media channels (television, radio, cell phone, fixed line
telephone, internet) to be disrupted following a strong local earthquake (directly or due to
loss of the electricity they rely on), these media should not be expected to be available to
provide informal warnings or confirm the occurrence of a tsunami. This increases the
importance of education to heed natural warnings.

Recommended guidance in New Zealand includes public response to informal warnings is


as follows, and appears on tsunami evacuation information boards:

GNS Science Report 2012/17 27


2012

“Warnings from friends, other members of the public, international media, etc. may be
correct. If you feel the threat is imminent, quickly get to high ground. Consider evacuating
from all zones. Verify the warning only once evacuated or if you can do so quickly via radio,
television, internet, or through your nearest civil defence emergency management office.”
(MCDEM, 2008a)

2.4 Formal warning messages and systems

2.4.1 Warning messages

The sequence of formal warning broadcasts from JMA in Iwate and Miyagi on March 11th
2011 is presented in Table 5. This sequence shows the significant increase in expected
wave heights (at the shore) in the first 41 minutes following the earthquake using the
framework discussed in Section 1.4. In addition to JMA warnings, there are some local
arrangements for providing early tsunami warnings. For example, in funato a tsunami
warning is disseminated by the disaster prevention office if shaking intensity registers
greater than 4 on their own intensity meter.

Although an early warning system is often seen publicly as key infrastructure in enhancing
tsunami resilience, the expectation of official warnings (and their content) may have slowed
the time taken for people to initiate evacuation in T hoku, compared to if there had been
total reliance on natural warnings. The broadcast of estimated tsunami heights had negative
consequences in Kamaishi, as the disaster prevention office continued to broadcast the JMA
estimate of 3 m even after JMA had released an updated estimate of 6 m, resulting in the
under-estimation of the tsunami threat by some of the public. In Kesennuma, the first waves
had already begun to arrive by the time updated wave heights were received by disaster
prevention officials, which somewhat negated the updates. In the case of Tar , expectation
that the 10 m high protective wall would protect the town from the expected 3 m high waves
may have slowed evacuation response times. Initial underestimation of the tsunami threat
was unavoidable due to the method used to estimate tsunami magnitude being reliant on
preliminary estimates of earthquake magnitude – preliminary earthquake magnitudes are
often revised upwards as more data becomes available. The difficulties in producing
accurate tsunami estimates from rapid estimations of earthquake magnitude were
recognised prior to this event and discussed by Imamura and Abe (2009).

Exposure to previous false ‘major tsunami’ warnings apparently led to complacency on


March 11th 2011, when wave heights commensurate with the ‘major tsunami’ warning
actually arrived. The evacuations of both Kesennuma and Natori were negatively affected by
complacency which resulted from a ‘major tsunami’ warning being issued following the
Maule, Chile earthquake of 2010. In that event, wave heights that occurred at Kesennuma
were 0.5-0.6 m and were approximately 0.5 m at Natori. The officials interviewed believe that
this led to subsequent underestimation by residents of the tsunami threat on March 11th
2011. This suggests some apparent confusion in public perception of anticipated tsunami
magnitude from local and distant tsunami sources. The threat of tsunami from the felt local
earthquake of March 11th 2011 was not perceived differently to the officially-warned distant
tsunami that had been expected from the 2010 Chile earthquake. Local officials do not
believe that an event two days prior, on March 9th 2011, contributed to this complacency –
despite the proximity of this event, the corresponding warning was for a lower level than on
March 11th.

GNS Science Report 2012/17 28


2012

Table 5 Time sequence of JMA tsunami warnings from March 11th to 13th, indicating the change
in estimated tsunami height for Iwate and Miyagi Prefectures. Between warning #3 and #10, the major
tsunami warning remained in place for these prefectures (Source: Japanese Meteorological Agency,
2011e).

Time Level for Iwate Level for Miyagi


th
March 11 , 14:49 (warning
Major tsunami: 3 m Major tsunami: 6 m
#1)
March 11th, 15:14 (#2) Major tsunami: 6 m Major tsunami: 10 m
th
March 11 , 15:30 (#3) Major tsunami: 10 m Major tsunami: 10 m

March 12th, 20:20 (#10) Downgraded to tsunami warning Downgraded to tsunami warning
Downgraded to tsunami Downgraded to tsunami
March 13th, 07:30
advisory advisory
March 13th, 17:58 Advisory cancellation Advisory cancellation

2.4.2 Warning systems

Japanese towns typically have extensive networks of speakers in the streets, which are used
daily for official announcements. In the event of a tsunami, these systems can be used to
broadcast a warning to the whole town, including spoken messages in addition to a siren
tone. The system often requires manual operation from a central broadcast point, such as
the city office. funato City had a broadcast system comprising over 120 mobile, portable
and fixed stations, some with visual alerts (Disaster Mitigation Section funato City, n d). As
is already widely recognised, warning systems are vulnerable to loss of power in local
earthquakes, an effect which was seen in Tar when broadcasts ceased shortly after the
audible warning broadcast began. Wireless transmission of warnings was also reported to
be ineffective in Natori City due to power outages.

Although the primary use of tsunami warning sirens is for warning of the tsunami risk, their
use in Japan for post-event welfare announcements or tsunami warnings in aftershocks puts
extra emphasis on their survivability, although further natural warning signs would be
expected in any large aftershock. Tsunamigenic aftershocks are a very real possibility
following a significant subduction zone event, as shown by the aftershock record around
T hoku which shows an offshore M7.3 event 3 months later on July 10th 2011 (Japanese
Meteorological Agency, 2011f). This is an interesting concept for further use of speakers and
siren towers. Scour, debris strike and inundation of electrical systems are all issues for
survivability of these networks – in funato, the systems failed when tsunami water
inundated battery power packs mounted on the siren towers.

A significant amount of information transfer in Minami-Sanriku is conducted by wireless


digital radio system. There were 105 speakers throughout the town and wireless radio
receivers are rented to every household by the municipal government. Public offices,
schools and factories also had wireless receivers (Figure 32), which cost JPY 50,000 (USD
640; NZD 850) per unit. They function on mains power with a battery in case of power
outage and are designed to remain on standby most of the time, automatically receiving the
tsunami warning. The tsunami warning was disseminated very effectively via this network on
March 11th 2011.

GNS Science Report 2012/17 29


2012

The original broadcast point of tsunami warnings in Minami-Sanriku was the disaster
management office, but due to severe tsunami damage to that building, a backup system at
the relocated town office was used for three days to announce welfare information and
aftershock warnings. This back-up system was only suitable for three days of use for welfare
announcements because the temporary facility did not have sufficient capacity for continued
operation. This illustrates the need for comparative redundancy in warning systems.

Figure 32 A wireless receiver shown by civil protection officials in Minami-Sanriku. These units are
rented by the local government to every household in the town, and resulted in effective warning
dissemination on March 11th 2011.

2.5 Movements during and after evacuation

Observations and reports of peoples’ movements during and after evacuation reveal that
many people died unnecessarily during the tsunami inundation, through social or parental
responsibility, lack of education or scepticism of warnings.

2.5.1 Delayed evacuation or non-evacuation

Delayed evacuation and non-evacuation were significant in influencing life loss of the
individual and of emergency responders or other community members. The importance of
immediate evacuation is highlighted by figures presented by Sagara (2011): only 1 person
died out of 33 who evacuated Rikuzentakata immediately. In contrast, 42 people died out of
147 people who did not evacuate immediately. Immediate evacuation increases chance of
an individual successfully evacuating. Emergency personnel such as fire-fighters, who have
responsibilities for specific tasks such as closing tsunami gates, may have very little time in
which to complete those tasks on behalf of the community. However, their lives were further
endangered when they remained in the hazard zone to mobilise people who had chosen not
to evacuate immediately.

As introduced in Section 2.3, older people in Natori City did not believe a tsunami would
affect the residential areas of Yuriage as they had no experience of tsunami reaching the
residential areas after previous local earthquakes. Others remained in the town to clear up
earthquake debris or to check on family members or neighbours, rather than evacuating.
Volunteer fire department personnel went to elderly peoples’ homes in order to persuade
residents to leave. This approach had some success as some residents decided to leave at
that point; however, as a result of their actions some volunteers were delayed from
evacuating the inundation zone and were killed in the tsunami. Natori City is a key example
of a location which suffered from a lack of recent experience, and underestimation of the

GNS Science Report 2012/17 30


2012

hazard through numerical modelling, which resulted in complacency and delays in


evacuation on March 11th 2011 despite the natural warnings. The issues caused by lack of
experience of local tsunami are likely to also affect Washington and New Zealand for the
same reasons, where emergency officials are working hard to prepare communities with no
previous experience of such an event.

Non-evacuation in Tar was reportedly caused by the communication of information that the
expected tsunami was only 3 m – this led to public perception that the town would be
protected by its tsunami defences, and some were sceptical that a tsunami would arrive at
all. Following the natural and formal warnings, some residents in Tar waited in the town for
family members to drive home from elsewhere in order to evacuate together. This had not
only the effect of delaying the person in Tar , but encouraging those in other locations to
drive into the hazard zone, with the result that some were unnecessarily killed during
inundation. The desire to evacuate with family was also reported in Kamaishi, where
residents are reported to have died while driving through the city having collected family
members.

In Kesennuma the disaster prevention official we interviewed believes that people should be
encouraged to help others in the stages of disaster preparation and post-evacuation, rather
than during the event when they should be concentrating on evacuating the hazard zone
themselves. In this way, people would be encouraged to not return or travel through
evacuation zones to help people prior to imminent tsunami arrival. He is considering
adopting the tsunami tendenko philosophy that was taught in Iwate Prefecture and attributed
to successful evacuations of schools in Kamaishi City.

These examples are important for Washington and New Zealand to consider, as delays in
evacuation are likely to cause many deaths in a local tsunami event, and education should
be put in place specifically to reduce this potential issue. Adoption in Washington and New
Zealand of a tendenko-like strategy encouraging self-evacuation and not stopping to help
others would be a significant change from current advice, which emphasises the importance
of immediate evacuation but does not explicitly advise “leave others behind” or “look after
only yourself”. In fact, advice from FEMA (2011) states the opposite, which may result in
further deaths from people remaining in the inundation zone too long:

“Remember to help your neighbors who may require special assistance - infants, elderly
people, and individuals with access or functional needs.”

Delays in evacuation were also caused by perceptions of safety close to the edge of the
recognised hazard zone. In funato City, most people in the official (and mapped) hazard
zones evacuated quickly, whereas those just outside the mapped hazard zone tended to not
evacuate until they saw the tsunami coming close to their immediate location. This effect
was seen in Minami-Sanriku, where neighbourhood fatality rates are higher in
neighbourhoods immediately inland of the mapped tsunami hazard zone, where people
perhaps perceived themselves to be safe. Murakami and Kashiwabara (2011) also suggest
this effect occurred in Natori City, where fatality rate in Yuriage 2 chome is 22%. In
comparison areas closer to the fishing harbour, suffered 11-12% a lower fatality rate
(Figure 33).

GNS Science Report 2012/17 31


2012

Figure 33 Building damage and fatality rate for seven areas of Yuriage, Natori City. Source:
(Murakami and Kashiwabara, 2011).

2.5.2 Use of motor vehicles during evacuation

Traffic congestion was a significant issue in several locations during the evacuation on
March 11th 2011, although other locations reported minimal congestion and attributed this to
previous guidance to evacuate the hazard zone by foot rather than motor vehicles.

In Tar , most people arrived at evacuation centres on foot, apparently following official
guidance to walk, rather than using motor vehicles. As a result, congestion was not a
problem in Tar . In Minami-Sanriku there was a limited amount of vehicle use in the
evacuation, which caused minor congestion initially but by 15:10 the roads were clear. Use
of vehicles during tsunami evacuation is prohibited in Minami-Sanriku due to (i) the
possibility of traffic congestion, (ii) failure of traffic signals during a power outage, (iii) poor
road conditions immediately following an earthquake, and (iv) evacuation locations may
quickly fill up with vehicles if they are used. During the annual evacuation drill, use of
vehicles is forbidden and despite the public here arguing that they should be able to use cars
for families and elderly to evacuate their homes, it remains prohibited.

The disaster prevention office of Kamaishi recognises that the importance of evacuating on
foot had not been impressed enough on residents. On March 11th traffic congestion occurred
on the Highway 45 Bridge at the port front, which forced some drivers to abandon their
vehicles and use the nearby government building as an ad-hoc vertical evacuation refuge. In
the February 2010 Chile tsunami, officials closed roads in the city to prevent people driving
into the hazard zone, but there were complaints from the public and people drove through
town anyway.

GNS Science Report 2012/17 32


2012

In order to combat traffic congestion in the town of Ryoishi, Kamaishi district, the town put in
place a scheme to designate certain vehicles and drivers to transport vulnerable people to
high ground in case of tsunami. The physical labelling of designated vehicles is aimed at
raising awareness of who is responsible for returning to the town and limiting the number of
vehicles being used in evacuation. On March 11th 2011 all designated drivers were out of
town at the time of the tsunami and the scheme could not be actioned before tsunami arrival.

Heavy use of vehicles in Ishinomaki led to road congestion and many deaths on March 11th,
despite repeated official messages at the time via loudspeaker not to use vehicles. Following
the April 7th 2011 aftershock, heavy use of vehicles occurred again, despite the impacts this
had during the previous evacuation.

The flat agricultural plains around Yuriage have no high ground for several kilometres and a
survey carried out in Natori determined that 65% of residents evacuated by motor vehicle
(Murakami and Kashiwabara, 2011). Many routes lead inland via a gridded network of roads,
suggesting many possible evacuation routes. However, these roads are all single-track and
deep ditches along road edges prevent two vehicles passing each other. For this reason,
residents of Yuriage chose to use main roads to evacuate in their vehicles. It is this type of
local knowledge that is invaluable in establishing suitable evacuation routes. The large
distances to safety across flat plains such as those surrounding Natori, provide a persistent
problem because people perceive that there is no other way to evacuate other than by car.
Murakami and Kashiwabara (2011) recommend the promotion of the use of bicycles to
reduce the amount of congestion in future tsunami evacuations.

A traffic accident occurred on the Yuriage Bridge – a major crossing point over the Natori
River – when a lorry shed its load, causing closure of one lane of traffic. This incident
illustrates the potential for unforeseen delays during vehicle-based evacuation, in addition to
traffic volume alone. Following a strong earthquake, falling debris or damage to roads or
bridges only increase the likelihood of such incidents. Experience of the Christchurch
earthquakes rendered many bridges unpassable for a short time and the potential for many
blocked roads following an earthquake in Wellington show that this is a very real problem for
major urban areas of New Zealand.

Education in Kesennuma has consistently stated that people should not use vehicles during
tsunami evacuation. However, this approach is under review following the March 11th event
in spite of traffic congestion prompting the fire department to instruct people to abandon their
cars. Due to there being some successful use of vehicles in this event and the recognition
that it is impossible to stop vehicle use entirely, the disaster prevention official stated that a
possible future approach in the city might be to accommodate vehicular transport by
widening roads or develop a road system that is designed specifically to cope better in an
evacuation situation. The installation of evacuation towers at junctions, or overhead
pedestrian walkways were suggested as potential vertical evacuation options in case of
traffic congestion and the need to abandon cars. This combination of developments is seen
by the authors as an unsuitable option, as this is likely to encourage people to use vehicles,
resulting in even more vehicles on the road and exacerbating current problems with
congestion. In addition the knowledge that evacuation towers are abundant may encourage
people to leave their vehicles at the earliest sign of congestion, thus blocking all vehicles
behind them. The authors believe it is more effective to emphasise evacuation on foot to
keep roads as empty as possible, as is current practice in Washington and New Zealand. In
recognising that some vehicles will be used, New Zealand advice is that if you have to use a

GNS Science Report 2012/17 33


2012

vehicle, drive as far as possible out of the evacuation zone to allow room for other people to
drive out, rather than stopping immediately outside the evacuation zone (MCDEM, 2010a).

2.5.3 Use of boats during evacuation

Residents of Hitachihamacho, Miyako City reported using fishing boats as a means of


escaping the tsunami and preventing damage to their boats. Many people sailed out of the
harbour at this location on feeling ground shaking, with the intention of sailing far enough
into deep water that the tsunami would not affect them. Once in deep water, a large group of
boats gathered together until the next day, as it was impossible to return to shore at night
through the large amounts of debris that had collected in the harbour. Some people travelled
out to sea in very small boats that had no means of shelter, and these boats were unsuitable
for the overnight weather conditions. It was reported that many other boats were still trying to
leave the harbour when strong currents of the initial waves prevented further progress. In
this location, deep water is only 1 km offshore, and the fishermen felt they would be safe
there, but it is clear that this approach is fraught with issues of precise timing before wave
arrival and debris fields, and is not a course of action to recommend to the general public
onshore. The International Tsunami Information Center advises evacuation by boat only in
the case of a distant tsunami (International Tsunami Information Center, 2006). New
Zealand tsunami advisories and warnings advise people to stay out of the water, including
any “boating activities” (MCDEM, 2010a), but there is no explicit mention of evacuation by
boat.

2.5.4 Schools evacuation

The interviews highlighted several examples of parents travelling unnecessarily to schools to


collect children following the tsunami warning, with fatal consequences. In Tar , parents are
required to collect their children from school in the event of a tsunami warning. The school
was not inundated on March 11th, but parents are said to have died while driving through the
town on their way to or from the school as tsunami waves arrived. Deaths of parents and
children were also reported in Kesennuma City during collection from schools. In Ishinomaki,
some parents collected children from schools and lost their lives while driving through an
inundated zone, whereas the children who remained at those schools survived on upper
floors of the buildings.

At Unosumai, Kamaishi City, all children successfully evacuated the Kamaishi Higashi
Middle School and Unosumai Elementary School, located 800 m inland from the pre-tsunami
shoreline following the earthquake. Tsunami evacuation training had been conducted in
Kamaishi schools since 2005 and 5-10 hours of annual class time was spent on learning
about the tsunami hazard (D. Harding and B. Harding, 2011; MSN Sankei News, 2011). Due
to the unexpected height of the tsunami, the children abandoned plans to stay on the 3rd
floor of their school building to evacuate uphill, and then had to relocate further uphill twice
more during the event. Their training had stressed that they should assess the situation as
they see it and be able to respond to changing events, and this most likely saved many lives
on March 11th. This is an excellent example of effective education in schools.

A recent study of schools disaster preparedness education in New Zealand showed that
there is no requirement for tsunami exercises in schools, and although most schools
understand their responsibility to care for their students until parents or an appropriate
guardian collect them, many schools (including in coastal locations) have not carried out

GNS Science Report 2012/17 34


2012

tsunami drills (Johnson, 2011). There is also no specific advice around the actions of parents
following an earthquake – i.e. collection of students from school, or organised rendezvous at
a specified safe location.

In the United States, FEMA advice states:

“If the school evacuation plan requires you to pick your children up from school or from
another location. Be aware telephone lines during a tsunami watch or warning may be
overloaded and routes to and from schools may be jammed.” (FEMA, 2011)

This advice makes no mention of the potential that parents may have to travel through
potential inundation zones to get to or from the school. Education in Washington encourages
parents to travel to the schools designated assembly point rather than collecting children
directly from school.

2.5.5 Necessary relocation from evacuation refuges

It has been reported in the media that over 100 evacuation centres were inundated on March
11th 2011 (Unknown, 2011a). In Kamaishi, nine of a total of 96 refuges were inundated. It
should be noted that these refuges were not vertical evacuation facilities designed to provide
refuge above the inundation height within the inundated area; rather they were evacuation
centres which had been incorrectly perceived to be outside of the tsunami hazard zone. In
Tar , evacuation centres had been located outside of the 1896 and 1933 inundation extent,
believed to be the maximum possible inundation (Figure 8). Unfortunately one of these
refuges was inundated, killing people who had travelled there.

The case of Unosumai has already been discussed, where evacuating school staff and
pupils were required to reassess their situation and evacuate to a different place according
to the evolving hazard. In this case, they evacuated further uphill several times (I. Matsuo,
pers. comm., October 20th 2011). In Ishinomaki, people left the Kadonowaki Elementary
School when the school building became inundated. They were able to exit to the rear of the
school and immediately climb to higher land, due to the location of the school. This school
later caught fire while inundated with tsunami water. The response of evacuees in these
cases shows that judging the evolving situation and being able to make further decisions is a
worthwhile education strategy, although this should not be at the expense of ensuring that
any evacuation centres are certainly outside of the maximum tsunami inundation, and any
vertical evacuation buildings are designed to remain safe in the worst tsunami.

2.5.6 Leaving evacuation refuges too early

In Tar , a town with high awareness of tsunami risk and a history of tsunami education,
many residents died on returning to the town a short time after arriving at a refuge, because
they had seen no wave eventuate. This occurred in spite of local government advice that
once a tsunami advisory or warning is issued, people should remain on high ground until that
advisory or warning is lifted. The disaster prevention official noted that there is no legal
recourse for not following the advice, so people are legally entitled to return to low-lying
areas at any time.

The Ishinomaki City disaster prevention official also reported deaths which occurred when
people had successfully evacuated to uphill refuges, only to leave the refuge when the JMA-
estimated time of wave arrival had passed with no apparent realisation of the tsunami. Some

GNS Science Report 2012/17 35


2012

people who had stayed at high ground while inundation occurred returned to low-lying areas
around 17:00, when tsunami waters receded, only for the fourth and fifth waves to inundate
parts of the city and kill many more people. A similar effect was observed during the event in
Yuriage, Natori City, where people who were in evacuation buildings left those buildings after
the first waves receded. Subsequent waves resulted in the deaths of many of these people.

The occurrence of deaths in this manner can be seen as an issue for education and
understanding of uncertainty in i) arrival times, ii) the time of first t wave versus the largest
waves, and iii) education of people in the fact that tsunami are a series of waves often
occurring over many hours.

In New Zealand, language broadcast in a National Warning: Tsunami Threat to New Zealand
is as follows, including a brief statement to clarify the uncertainty of arrival time:

“If a tsunami has been generated, the first wave may arrive in New Zealand in the areas
around [insert place] at [insert NZDT/NZST on insert date]. The first wave may arrive later
and may not be the largest” (MCDEM, 2010a)

Recommended text for warning and evacuation information covers the other issues, stating:

“There may be multiple waves separated by up to an hour, or more. Large waves may come
after a series of smaller waves. The largest waves from distant sources may take many
hours to arrive.” (MCDEM, 2008a)

“Stay OUT of evacuation zones until given the official ‘all-clear’ from Civil Defence
Emergency Management. Stay away from the Red Zone for at least 24 hours after any
tsunami warning – even small waves can create dangerous currents.” (MCDEM, 2008a)

In Washington, advice in evacuation brochures give consistent advice to wait for an ‘all-clear’
message before returning to the shore. Washington Military Department Emergency
Management Division (1999) states:

“Do not return to shore after the first wave. Wait for Emergency Management officials to give
the “All Clear” before you return.”

A later map for Ocean Shores published by Washington State Department of Natural
Resources covers several points in this statement:

“The first wave is often not the largest; successive waves may be spaced many minutes
apart and continue to arrive for several hours. Return only after emergency officials say it is
safe” (Washington State Department of Natural Resources, 2007)

2.6 Impacts of coastal defence on warning response and evacuation

Japan spends a significant amount of resource on ‘hard’ coastal defence, and some of the
most formidable sea walls and breakwaters of the T hoku region were adversely affected
during the tsunami of March 11th. It is recognised that these substantial defences can
prevent damage in small to moderate events, and may allow extra time for evacuation in
such extreme tsunami. The breakwaters at Kamaishi are estimated to have reduced
inundation height by 40%, and delayed onshore arrival by approximately 6 minutes
(Takahashi et al., 2011). Despite the physical benefits, our interviews also indicate some

GNS Science Report 2012/17 36


2012

negative impacts of significant hard defences, which can induce false impressions of safety
and an inability to see the natural warning of a disturbed ocean. Even in small events,
unforeseen issues (e.g. break-down and non-closure of gate systems/earthquake damage)
may result in unexpected inundation.

The tsunami impacts in Tar have been widely reported due to the damage the town
sustained, despite having a concrete sea wall of 10 m height, in three sections with a
combined length of 2,433 m (Tar Town Municipal Government, n d). This sea wall
prevented damage in the town due to the 1960 distant Chile tsunami, and would likely
prevent damage to the main town in the vast majority of tsunami in its design lifetime.
However, on March 11th, the perceived safety in which residents lived may have contributed
to a number of deaths – as already discussed in Section 2.5, the presence of the wall and
the estimated 3 m tsunami height combined such that some people believed they would be
unaffected behind the wall. Sixteen people died while driving through Tar , possibly as a
result of not feeling the earthquake or at least not being able to sufficiently recognise the
arriving tsunami due to their views of the sea being obstructed. The disaster prevention
official also reported that some people believed that the tsunami spray hitting the wall was
fire rather than waves arriving.

This issue was also apparent in Ryoishi, where several vehicles were observed driving along
the coastal road immediately adjacent to a sea wall (Bombaadi, 2011). At the time these
vehicles were on the road, the tsunami was rising rapidly on the other side of the wall
unseen by the drivers before it poured over the top of the wall, which was several metres tall.
The reduction in visibility out to sea also endangered the lives of fire-fighters in Setai, Miyako
City, who were manually closing tsunami gates when they saw the tsunami at the last minute
through a small window in the gate and were able to drive away.

New Zealand and Washington presently have sea walls (mostly around harbours) designed
only to mitigate normal storm waves and prevent long-term erosion. On coastlines at risk of
a local subduction zone earthquake, it is unlikely to be feasible to construct a defence of
sufficient size to protect against the maximum potential events – exactly as shown in Tar .
Sea walls may reduce tsunami velocity and wave heights even if breached or damaged,
however, the population of the area ‘protected’ by the defence must be suitably educated to
know that the defence is not a 100% guarantee of safety.

2.7 Vertical evacuation buildings

Vertical evacuation buildings are designed to provide refuge in the inundated area by
providing sufficient elevation above the maximum water level. This type of refuge proved to
be extremely valuable in reducing life loss on March 11th 2011. These buildings are different
to regular evacuation centres, which are generally designated in Japan for use in multiple
hazards, and are located outside of the previously estimated tsunami hazard zone.

The consistent education message delivered to the public in T hoku was to evacuate inland
or to high ground as priority, but the presence of vertical evacuation buildings was valuable
in situations where people were unable to get inland or to high ground prior to tsunami
arrival. The successes are numerous; however, some important lessons can also be learnt
from the experiences of evacuees who used these buildings. Appendix 4 provides a list of
vertical evacuation buildings in the locations visited, with details on construction, signage,
access, inundation level on 11th March and number of people saved.

GNS Science Report 2012/17 37


2012

Section 2.1 introduces the Project Safe Haven initiative underway in Washington to develop
vertical evacuation strategies. Evacuation buildings are not currently designated in New
Zealand, but some existing buildings could be considered suitable. An initial scoping study
was recently completed and the Ph.D. thesis work of Stuart Fraser is focussing on
developing guidelines for vertical evacuation in New Zealand; this report forms part of the
basis of that research.

2.7.1 Building specification and levels of damage

The requirements of buildings to be designated for vertical evacuation by local government


in Japan are that they meet adequate construction standards and minimum heights as
specified by Hiroi et al. (2005). In consideration of local sources, these buildings must
conform to sufficiently high seismic standards to provide life safety following a significant
local earthquake. The seismic safety requirement of vertical evacuation buildings in Japan
was achieved by designating buildings constructed to post-1981 seismic standards.
Although some of the vertical evacuation buildings observed suffered damage from ground
shaking, none suffered sufficient damage to prevent use in the subsequent tsunami
evacuation. Detailed structural analysis is out of the scope of this report.

The key criteria from the 2005 guideline for official designation of buildings as tsunami
evacuation buildings are that they:

• Are of a minimum height according to estimated maximum inundation depth


̇ Less than 1 m depth = 2-storeys or higher required
̇ 2 m depth = 3-storeys or higher required
̇ 3 m depth = 4-storeys or higher required
• Are reinforced concrete (RC) or steel reinforced concrete composite (SRC) construction
• Were constructed after 1981 (the latest significant update of building codes in Japan)

Building damage surveys following this event underline previous observations that reinforced
concrete structures are the most effective buildings in tsunami, as they are the most
resistant to tsunami loading, and debris impact. Structural members of observed steel
buildings generally survived tsunami loading very effectively but high levels of damage to
cladding and other non-structural elements (EEFIT, 2011), and high potential for significant
damage from debris strike makes them unsuitable for use as a safe refuge.

There are cases of some vertical evacuation buildings being overtopped (e.g. Fisheries
Cooperative in Minami-Sanriku) and in other cases the maximum inundation height was
close to the building roof when the tsunami arrived at low tide. This has prompted some of
the interviewed officials to call for a minimum height of 5-storeys for designated facilities.
Despite this, there were some 2-storey buildings that proved sufficient as a refuge in places
where flow depths were relatively low in 2011 (e.g. Yuriage).

2.7.2 Observed examples of vertical evacuation buildings

There were 3 designated vertical evacuation buildings in Kamaishi district at the time of this
event: an 8-storey RC frame apartment block and a 2-storey RC government building
(Figures 34, 35 and 36) in the city, and a 4-storey steel frame hotel near Unosumai (Figure
37). An estimated 50 people evacuated to the apartment block and government building.

GNS Science Report 2012/17 38


2012

Damage was limited to glazing and very minor cladding damage on the government building.
Damage to glazing and non-structural RC-infill panels was common at the apartment block,
which was inundated to the 3rd storey, while debris impact had damaged a steel-frame
vehicle parking elevator (EEFIT, 2011). People were trapped in this building until March 12th
due to silt blocking the stairwells (EEFIT, 2011). The owner of the Hotel Horaikan at
Unosumai directed guests to high ground behind the hotel, rather than staying in the
building. The hotel suffered non-structural damage at the ground floor only, and at the time
of our visit the hotel was being refurbished internally.

There were plans for further tsunami evacuation buildings and towers close to the Kamaishi
Port, but these had not yet been constructed.

Figure 34 Designated evacuation locations in Kamaishi City. An estimated 50 people evacuated to


this apartment block (A) and government offices (B). The entire area shown was inundated;
inundation depth was 8 m in the vicinity of building (A).

Figure 35 Apartment block in Kamaishi City, which was inundated to the 3rd storey
Figure 36 Government offices in Kamaishi City. This building is raised several metres above sea
level and sustained damage to only the 1st storey windows

GNS Science Report 2012/17 39


2012

Figure 37 The Hotel Horaikan at Unosumai, Kamaishi City, with official signage displayed at the 3rd
floor and very little external damage. The hotel was inundated to the ceiling of the 2nd storey

Figure 38 Locations of 5 designated vertical evacuation location visited in the downtown area of
funato City: (A) Plaza Hotel; (B) Yasuka office/apartment buildings; (C) Maiya Shopping Centre and
car park.

Seven buildings in funato City were recognised by the community and municipal
government as vertical evacuation buildings, five of which we visited during our field visit
(Figure 38). These were not officially designated as they did not meet the standards set by
the 2005 government guidelines due to their construction prior to 1981. As result, the
buildings are defined through informal arrangements with the building owners. This
arrangement is in place for the Maiya shopping centre and adjacent multi-storey car park
(Figure 39 and 40), Yasuka private commercial and apartment blocks (Figures 41 and 42),
Plaza Hotel (Figure 43 and 44), Fukutomi Hotel, Kesennuma Banking building, Nokyo Credit
Co-operative building, and the Wedding Plaza. Despite the construction of these buildings
prior to latest seismic requirements, they were not significantly damaged in this earthquake.
The only structure less than 4-storeys in height is the RC construction Wedding Plaza, which
was inundated almost to its roof but survived intact. Inundation height in the vicinity of these
buildings was around 3-storeys (9-10 m height; EEFIT, 2011).

GNS Science Report 2012/17 40


2012

Figure 39 Maiya shopping centre, which was inundated to the 3rd storey
Figure 40 Maiya multi-storey car park showing damaged railing at the 2nd storey.

Figure 41 Yasuka private commercial and apartment block number 1 shows very little exterior
damage.
Figure 42 Yasuka private commercial and apartment block number 2 is adjacent to block number 1
and was being refurbished at the time of our visit

Figure 43 The Plaza Hotel, which was being refurbished at the time of our visit
Figure 44 External stairs giving access to the 4th storey of the Plaza Hotel

Twelve buildings were designated for vertical evacuation by Kesennuma City in 1982 with a
further three added later. Nine of these buildings are shown in Figure 45, and individual
photographs are shown in Figures 46 to 53. These buildings provided refuge to 2,326 people
on March 11th. With the exception of the car park deck over the fish market (which was
constructed with vertical evacuation as a planned function) these buildings were all existing

GNS Science Report 2012/17 41


2012

structures that met the 2005 guidelines. Being constructed over a fish market, parts of the
car park structure are open with steel columns and no walls, enabling flow through the
structure.

Consideration is being given to changing the requirements for buildings used for vertical
evacuation in Kesennuma, as several buildings were close to being overtopped. The
tsunami arrived at low tide, and it is estimated that overtopping of vertical evacuation
buildings would have occurred if the tide and tsunami combined had been just 1 m higher. In
recognition of this, the fire department believes that buildings designated for vertical
evacuation in future should be 5-storeys or greater. Additional features under consideration
include night-time lighting to indicate evacuation routes, and emergency power supplies in
vertical evacuation buildings. Fires were a significant issue in Kesennuma, after 51,000 litres
of oil was spilt from ruptured oil tanks. Many buildings burned and some evacuation centres
narrowly avoided catching fire while occupied by evacuees. A government committee has
been set up to prevent such a spillage of oil occurring again, and the fire-proofing of
evacuation structures was raised as a consideration for future design of such facilities.

Scour of vertical evacuation buildings was common in Kesennuma and was observed at the
office building (Figure 46), welfare centre (Figure 48) and Prefectural Office (Figure 50). The
buildings were generally undamaged by debris strike despite their location in a busy port,
with damage from debris only observed at the office building in Figure 46.

Figure 45 Locations of designated tsunami evacuation buildings in Kesennuma City (Data from
Kesennuma City General Affairs Department). (A) Office building; (B) Yoyoi food factory; (C) Shoe
store; (D) Elderly welfare centre; (E) Tourist Pier and fish market; (F) Prefectural Office; (G) National
government office; (H) Central Community Centre; (I) Kahoku Newspaper office

GNS Science Report 2012/17 42


2012

Figure 46 An office building in the inner harbour was unused on March 11th. This building sustained
debris damage to the external stairs (front right) which face the harbour. Minor scour occurred along
the seaward face of the building, and around the base of the external stairs.

Figure 47 The Yoyoi food factory, where 400 people took refuge. Inundation height was around 8 m
here.

Figure 48 An elderly persons’ welfare centre which was inundated to around 8 m height. Scour up
to 2 m deep occurred on the near side of the building in this photograph. Extensive damage occurred
immediately inland of this building, while 80 survived on the upper floors.

GNS Science Report 2012/17 43


2012

Figure 49 The large open car park roof space above the fish market provided refuge to 1000
people. This structure is at the harbour front and inundation almost reached roof level. The open steel
structure enabled tsunami flow at the 1st storey.

Figure 50 The Prefectural Government offices suffered extensive scour and heavy in inundation of
8 m. This building was constructed after 2004 and displays tsunami vertical evacuation signage at
roof level. 200 people survived the tsunami in this building

Figure 51 The National Government office, where 120 people took refuge from the tsunami. This
building is adjacent to the Prefectural Government offices and was left undamaged when light steel
structures were washed into it. The lower two storeys were inundated.

GNS Science Report 2012/17 44


2012

Figure 52 Kesennuma central community centre, where 450 people took refuge when they
evacuated nearby buildings
Figure 53 The Kahoku Newspaper offices were inundated to at least the ceiling of the 1st storey. 71
people took refuge here.

There were four designated tsunami evacuation buildings in Minami-Sanriku in the near
shore area (Figure 54) which was inundated to 11 m above ground level:

• The Matsubara community apartment block (Figure 55), which is owned by the town, is
located at the harbour front. It is RC construction, 4-storeys in height, and 44 people
survived on the roof, despite inundation to the 4th storey. This building was constructed in
2007 and planned as a vertical evacuation structure to provide refuge to large crowds at
the adjacent sports ground. This is the only building in Minami-Sanriku with tsunami
evacuation signage, due to its construction after use of common signage began in 2004.
Due to its location, there were initial public concerns about the building being used for
tsunami evacuation, but it was agreed the roof level would constitute safe elevation.
Significant scour occurred around this building, with at least 2 m of scour around the
foundations (EEFIT, 2011).

• The Takano Kaikan wedding ceremony building (Figure 56) provided refuge to 330
people, most of whom were elderly and attending an assembly at the time of the
earthquake. The building owners considered use of this building in evacuation as
corporate social responsibility.

• Shizugawa Hospital is a site comprising two buildings of 4- and 5-storeys in height


(Figures 57 and 58). Most of the 320 people who were in, or evacuated to this building
prior to the tsunami survived on the roof but 71 people were killed, the majority of whom
were patients who could not relocate to the top storey due to illness/weakness.

• Despite being only 2-storeys in height, the fishing co-operative building (Figure 59) was
designated for vertical evacuation use through its ownership by a public organisation
looking to protect its workers. The building was unused on March 11th.

GNS Science Report 2012/17 45


2012

Following experiences of March 11th, the disaster prevention official in Minami-Sanriku


recommends that all vertical evacuation buildings should be higher than 5-storeys.
Evacuation towers may be considered in the town for the future but as was the case before
March 11th, primary advice will be to evacuate to higher ground rather using buildings or
towers. These should remain a last resort due to the potential for people to become isolated
in buildings and require rescue later.

Figure 54 Locations of designated tsunami evacuation buildings in Minami-Sanriku Town: (A)


Matsubara apartments; (B) Takano Kaikan building; (C) Shizugawa Hospital; (D) Fisheries Co-
operative. Significant scour can be seen where seawater is encroaching around building (A) and to
the east of building (D).

Figure 55 Matsubara apartment block viewed from the seaward side with waterborne debris to the
4th storey. The low seawall has been destroyed in the tsunami return flow and was sheltered in part by
the apartment building. The foundations of this building are exposed above the water level due to
extensive scour.

GNS Science Report 2012/17 46


2012

Figure 56 The Takano Kaikan building 160 m inland of the Matsubara apartment block. Little
damage aside from glazing was observed here, despite heavy damage and collapse of RC buildings
immediately seaward.

Figure 57 Shizugawa Hospital (west building) – the taller of the 2 hospital buildings at 5-storeys,
this section suffered less debris damage than the east building.
Figure 58 Shizugawa Hospital (east building) viewed from the seaward side which suffered debris
damage at the 1st and 2nd storeys.

Figure 59 Minami-Sanriku fisheries co-operative building which was overtopped on March 11th.
Image credit: Google Streetview, August 2011.

GNS Science Report 2012/17 47


2012

There was widespread use of buildings for informal (unplanned) vertical evacuation in
Ishinomaki on March 11th, 2011. Following production of the 2005 government guidelines on
vertical evacuation, Ishinomaki City office reached agreement with three private companies
in the Minato-machi district of the city to use their facilities for evacuation. On March 11th, a
total of around 500 people sought refuge at these three buildings: York Benimaru shopping
centre (Figure 60), Homac hardware centre (Figure 61), and Hotaru funeral facility (Figure
62). Although these buildings are only 2-storeys in height, they were deemed appropriate for
the city’s level of tsunami risk and were sufficient for the tsunami inundation which occurred.
Two of the buildings (Homac and Hotaru) have external vehicle ramps leading to rooftop car
parking areas; the funeral facility appeared not to have external access to the 2nd storey, and
it is unclear whether 24 hour access is available.

There was no disagreement from the building owners when they were approached by the
city. Once the use of these buildings had been agreed, the agreement was broadcast on
local news but their function was not publicised widely and no signage was applied. It is
agreed that the city will pay compensation to the building owners in the event of damage or
costs incurred to the property when people evacuate to the property – for example, people
having to break in or occupy the building for long periods of time.

In addition to these three designated buildings, almost any building that is higher than a 2-
storey residential structure was used for vertical evacuation in this event. About 260 official
and unofficial evacuation places were used in total, providing refuge to around 50,000
people. These included schools, temples, shopping centres and housing. In addition to this,
there were another 50,000 people trapped in the upper storeys of houses. Due to the effects
of this tsunami, it is recognised that more evacuation structures are required west of the
Kitakami River. Relocation of evacuees was required in some locations, for example at
Kadonowaki Elementary School (Figure 63), which was inundated at the first floor and was
subsequently affected by fire within 1 hour of the earthquake. Those people that had
evacuated to the school were able to leave the school prior to the fires, moving further inland
and uphill to another school.

Figure 60 York Benimaru shopping centre, Minato-machi, Ishinomaki City. The rooftop car parking
area is accessible by vehicle ramp and was successfully used for vertical evacuation on March 11th.
Figure 61 The Homac hardware store is a single storey building with car parking on the roof, which
is accessible by vehicle ramp. Image credit: Google Streetview, August 2011

GNS Science Report 2012/17 48


2012

Figure 62 The Hotaru funeral facility. This is an unlikely vertical evacuation facility given the lack of
obvious roof space or external access, but successfully provided refuge on March 11th.
Figure 63 Kadonowaki Elementary School was damaged by fire while inundated during the
tsunami. People who had sought refuge hear were able to leave the building and evacuate further
uphill at the rear of the school. This building is a typical school structure in Japan; these buildings are
commonly used as evacuation or emergency welfare centres.

Four tsunami buildings in Natori City had been specified by the municipality government as
evacuation places: Yuriage Community Centre (Figure 65), Yuriage Junior High School
(Figure 66), and Yuriage Elementary School (Figure 67). These buildings are situated
outside of the previously estimated tsunami hazard zone, so it is believed that these
buildings had been identified as regular evacuation centres, rather than specifically for
vertical evacuation. Sendai International Airport in Kitakama was also an evacuation
location, through an existing agreement with local residents that it would be used in the
event of a tsunami. These buildings were the destination for many people evacuating
Yuriage (Murakami and Kashiwabara, 2011) and a total of 3,285 people evacuated to these
buildings (Unknown, 2011b), but other 1 to 2-storey buildings between Yuriage and the
airport, such as a pump-house and a university boat club building were also successfully
used for refuge. People also survived on a pedestrian footbridge close to the Yuriage Bridge.
The designated schools in Yuriage are 3-storeys in height and the community centre is 2-
storey. In this event a small amount of water reached the 2nd storey but people survived at
that level.

GNS Science Report 2012/17 49


2012

Figure 64 Locations of designated evacuation buildings in Yuriage, Natori City: (A) Yuriage Junior
High School; (B) Yuriage Elementary School; (C) Yuriage Community Centre. Sendai International
airport is to the south of the area shown here.

Figure 65 The Yuriage Community Centre was inundated to the 2nd storey but 43 people survived
in the building.
Figure 66 Yuriage Elementary School. 870 people took refuge here, where the large roof area and
3rd storey were above inundation. Direct external access to the 3rd storey is available from the stairs
shown here.

GNS Science Report 2012/17 50


2012

Figure 67 Yuriage Junior High School is constructed on ground which is raised 1.8 m above the
surrounding fields. This helped to mitigate damage from the tsunami flow and only contents damage
was observed (EEFIT, 2011). 823 people took refuge here.

2.7.3 Community engagement

Community-led identification of buildings that may be suitable for vertical evacuation was
common in the locations visited, particularly in funato City where all vertical evacuation
buildings were originally proposed by the community, but did not meet the age requirements
set by the government guidelines. In Kamaishi, the community identified several buildings
that they believed could be used in a tsunami, additional to the three officially designated
buildings. The suggested buildings did not meet the government requirements and the local
authorities do not advise their use during tsunami. Only the building owners used these
buildings on March 11th 2011, and although the lowest 3-storeys were damaged, the
occupants survived.

Community disaster prevention groups in Kesennuma approached the owners of the Yoyoi
food factory in the Hamacho neighbourhood about using it for vertical evacuation, after
which the building owners and the city came to an agreement to designate it as a vertical
evacuation building. The building has signage over the doors and is used in ongoing training
as part of the evacuation drill held at the factory twice a year with the local fire department
and residents.

2.7.4 Evacuation building signage

Consistent vertical evacuation signage for buildings was introduced in 2004 and in general
only buildings constructed after 2004 have this signage in place on their exterior (e.g.
Figures 68 and 69). The only observed buildings with official signage in place are the
apartment block in Kamaishi; Hotel Horaikan in Unosumai; Matsubara apartment block in
Minami-Sanriku; the Prefectural Office in Kesennuma City; and Yoyoi food factory in
Kesennuma.

GNS Science Report 2012/17 51


2012

Figure 68 Official vertical evacuation signage, as photographed on the apartment block in


Kamaishi.
Figure 69 Sign indicating the entrance to external stairways at the apartment block in Kamaishi
(Translation: “Evacuation building entrance (stairways)”.

2.7.5 Owner agreement

During designation of buildings, disaster prevention officials found building owners to be


“extremely cooperative” in Kesennuma, and this is also seen in other locations where
interviews were carried out. The owner of the Hotel Horaikan in Kamaishi district proposed
that her own building be designated as an alternative option to defences blocking beach
access. She had previous experience of evacuation into a building following a tsunami
warning, so built her hotel as a 3-storey building so it could be used in tsunami. The owners
of the Takano Kaikan building in Minami-Sanriku were said by disaster prevention officials to
recognise the corporate social responsibility of agreeing to this use. The Minami-Sanriku
fisheries co-operative was interested in designation to protect its workers.

In funato, the disaster prevention official reported some initial resistance from building
owners when they were first approached about the potential use of buildings in evacuation.
The owners’ concerns focussed on night-time access and who would be responsible for
evacuees while in the building, but following discussions with the community the owners
agreed to their buildings being used. We were unable to confirm through our interviews, the
reasons for final agreement. The issue of responsibility for evacuees remains a key issue
here: employees at the Credit Co-operative chose to leave the building on hearing the
tsunami warning on March 11th, and it is unknown what impact this had on people trying to
access the building in this case.

2.7.6 24-hour access

A vital issue in the effective use of vertical evacuation buildings is guaranteed access 24
hours a day, on any day of the year. Access to upper floors of buildings in Japan is often
provided via external stairs, which exist on many buildings for emergency egress in fire or
earthquakes. These stairs may not always lead to the roof, but do at least give access to
upper floors. Where access is not available by external staircases, several other methods of
access are available in the locations visited (see also Appendix 4):

• Owners of private buildings have security staff present overnight to allow people in (e.g.
Prefectural Office in Kesennuma)

GNS Science Report 2012/17 52


2012

• Building owners agree to breakage of doors and windows to allow emergency access
(e.g. Kesennuma Junior High School and National Office in Kesennuma)
• Due to day to day use, the building is staffed, or has residents in a night-time 24-hrs a
day (e.g. Shizugawa Hospital, Hotel Horaikan, Matsubara apartments)
• Representatives of local residents have keys to enable access outside of office hours
(e.g. two representatives living near to government building, Kamaishi; Yuriage schools
and community centre, where the key holders are informed by telephone that they need
to open the facility).

2.7.7 Evacuee Welfare

Welfare of evacuees is another important issue in considering use of vertical evacuation


buildings. Evacuees had been stranded in many buildings for up to 2 days following the
earthquake, due to remaining tsunami waters or debris blocking egress from the building.
Welfare considerations are primarily provision of food and water, shelter, warm blankets or
clothing, sanitation facilities and a means of emergency communications with disaster
prevention headquarters or emergency services.

Provisions were available at the Prefectural Office in Kesennuma and South Kesennuma
Elementary School but these were appropriate to a 6 hour occupancy period only. It had
been assumed that after 6 hours residents would be able to get to welfare centres where
more long-terms provisions are stored prepared in the event of a disaster. However, some
residents had to remain in these buildings until the March 13th, when they were either
rescued by helicopter or by road once debris had been cleared. Other examples of evacuees
having to remain in evacuation buildings until March 12th were at Yuriage Elementary and
Junior High schools, and the apartment block in Kamaishi, when evacuees had to leave
through tsunami water or dig silt out of stairwells to exit the building. Evacuees remained in
the airport terminal until March 13th until they were rescued. The Yuriage community centre
and Junior High School both had some provisions for evacuees, but there was very little at
the school and storage of provisions in the community centre was on the ground level, which
became inundated. There was no emergency communications equipment in the schools.

In funato, provisions were available at the shopping centre due to the day-to-day function
of the building, but no specific arrangements had been made to provide short-term support
for occupants in the event of a tsunami. Apartment buildings are also likely to have some
provisions and shelter, given the regular residential use of the building. The disaster
prevention official here raised the provision of a communications link in all refuges as an
important resource, to enable evacuees to contact help in case of requiring urgent rescue
(for example if threatened by fire or serious illness) when cell phones or other radio systems
are not functioning.

The disaster prevention official in Minami-Sanriku re-affirmed an important challenge when


planning resources for evacuation buildings: it is difficult to say how many people will use
any given building. More work is required on resource planning, but it is clear from the
events of March 11th that basic resources are highly desirable for a few days of occupancy.
However, this issue is secondary to the provision of life-safety, and designation of suitable
buildings should not be delayed or prevented as a result of inadequate welfare facilities.

GNS Science Report 2012/17 53


2012

3.0 RECOMMENDATIONS

This section presents recommendations for Washington and New Zealand, as a result of the
authors’ observations and information from interviews. These recommendations refer to the
themes discussed previously in Section 2.0 under corresponding headings.

3.1 Community involvement in evacuation planning and preparedness

There is well-developed community engagement in tsunami preparedness in T hoku. This is


extremely positive for perpetuating community awareness of hazard and appropriate
response, and increasing community ownership of disaster planning. In the municipalities we
visited, 96% of people living in the inundated area survived. This is a very positive result in
the context of such a large magnitude earthquake and tsunami, and the high rate of survival
can be attributed to this high awareness and subsequent effective evacuation.

Community engagement should continue to be encouraged as it currently is in Washington


and New Zealand, through projects emphasising community input and ownership.

3.2 Hazard/evacuation maps and signs

3.2.1 Maps

The maps provided to coastal residents in T hoku show some of the same inconsistency
issues recently highlighted in the Pacific Northwest by Kurowski et al. (2011). More focussed
research of map content and style (and their impact on evacuation response in T hoku) has
the potential to provide recommendations of regional scale evacuation mapping best
practice. This can benefit programs that are already under way in Washington and New
Zealand to enhance tsunami evacuation mapping standards. Researchers should monitor
any Japanese research on the impacts of hazard and evacuation mapping styles and
inconsistencies had on evacuation response on March 11th 2011 and assimilate research
findings into current efforts to improve consistency.

It is encouraging that the U.S. and New Zealand have already taken steps to improve
consistency and clarity of their mapping. At this time when Japan needs to revisit its
mapping, there should also be opportunity U.S. and New Zealand guidelines to positively
inform the next generation of maps in Japan.

Showing evacuation shelters and/or assembly areas on maps is recommended but maps
should only show those sites which have been officially designated as being at sufficient
elevation to ensure safety in tsunami, and of sufficiently strong construction to survive a local
earthquake and tsunami. Evacuation routes should be shown to encourage people to
evacuate out of the evacuation zone as soon as possible and then to shelters, rather than
travelling extensively within the evacuation zone. Confusion among evacuees may arise
where shelters or assembly areas are designated for use in the occurrence of different
hazard events (e.g. landslide), and may be located within the tsunami evacuation zone.
These should not be included on tsunami evacuation maps, and there should be clear,
specific education informing the community about whether or not that facility is suitable for
use in tsunami evacuation.

GNS Science Report 2012/17 54


2012

3.2.2 Hazard/evacuation signs

Washington and New Zealand already have well-developed tsunami signage guidelines and
our observation of similar signage in Japan suggest only a minor addition. It is
recommended that consideration be given to including on signs the distance to safe location,
in order to provide some perspective of the distances people would be required to cover in
an event.

3.3 Response to natural and informal tsunami warnings

Education that people should evacuate coastal areas if they experience a strong or long
earthquake was an appropriate concept for T hoku and the benefits of this education were
shown on March 11th 2011 with high proportions of the population evacuating immediately.
This is an example that the message currently given in New Zealand is the correct one for
local tsunami, and should continue to be used.

Relevant agencies should continue education which encourages immediate evacuation upon
experiencing natural warning signs. Particularly encouraging is the fact that immediate
evacuation can have a cascading effect on influencing others (who may be disoriented or
unaware of appropriate action) to evacuate.

An important point to note is that personal or community experiences appeared to negatively


affect the proportion of people evacuating immediately. Due lack of experience of local
tsunami in Washington and New Zealand education must be made more effective to counter
this and achieve high proportions of immediate evacuation.

In the United States, preparedness advice from relevant agencies on the West coast varies
but maintains the core message to evacuate on feeling ground shaking. This message is
unfortunately not replicated by the message given by FEMA (2011), whose current advice is
to listen to the radio upon feeling an earthquake. A subsequent FEMA statement advises
immediate evacuation upon receiving an official warning. Even if the radio is still working
during the earthquake, strong or long ground shaking is by far the fastest warning and
advising people to monitor the radio will likely (a) slow evacuation and (b) reduce public
confidence in their ability to act on natural warnings in the absence of official warnings.
Given the short expected arrival times (20-30 mins) of the first waves at the Washington
coast, the FEMA advice should be made consistent with that provided by Washington EMD
and WCATWC, to evacuate to high ground upon experiencing a strong earthquake.

It is recommended that additional clarification be added to provide a clear statement of what


people can expect to experience in a strong earthquake – i.e. experiencing difficulty in
standing up. Additional consideration should also be given to the occurrence of long periods
of sustained ground shaking (longer than a minute), when people may experience lower
levels of shaking that do not trigger response to the earthquake strength.

Education in New Zealand states that informal warnings from friends and members of the
public may be correct and that people should not wait but rather evacuate and then confirm
the warning once at a safe location. Informal warnings were valuable in Japan and
discussion supporting informal warnings should be included alongside any education around
formal and natural warnings.

GNS Science Report 2012/17 55


2012

Communications channels can be disrupted in a local earthquake event, and the population
should be encouraged to recognise the unreliability of these communication channels in
such a situation and respond to the natural (or informal) warning without scepticism or
hesitation.

3.4 Formal warning messages and systems

3.4.1 Warning messages

In general there is a high expectation in Japan that official warnings will be provided for local
tsunami, in addition to regional or distant tsunami. In Washington and New Zealand there is
a much longer response time for the broadcast of official warnings, precluding their use in
local events. In addition, the high monetary cost of implementing technology required for a
Japanese-style early warning system means that at present, more effective evacuation of
the population in local source events can be achieved in Washington and New Zealand if
people are educated to evacuate independently of formal warning (i.e. in response to natural
warnings).

At a time when official hardware-based warning systems are growing in scale and apparent
technological advancement, there is increased potential for people to be less likely to
respond to natural warnings. While recognising the benefits of tsunami warnings for regional
and distant tsunami, there should be continued investment into education of the different
response issues for local tsunami versus regional and distant tsunami.

Reliance on early warnings for local source events in Japan has previously caused frequent
false alarms (and unnecessary evacuation), which in some areas led to complacency and
meant some people delayed or declined to evacuate on March 11th. By comparison,
earthquakes that are long (lasting more than a minute) or strong (enough that people can’t
stand up) are rarely felt in a person’s lifetime in Washington and New Zealand and any use
of official warnings should not detract from public response to this unmistakable natural
warning. Education material and especially community planning and exercises that clearly
identify the difference in size between expected distant source and local source tsunami are
required to minimise normalisation bias that occurs from experience of more frequent ‘small’
distant tsunami.

The use of broadcasts explicitly stating expected wave heights should be considered
extremely carefully before implementation. Miscommunication occurred in T hoku when
communications systems failed. Inaccurate estimates and failure in the system was shown
to result from technological failure, human error, and due to reliance on preliminary
earthquake parameters to estimate tsunami height. A more conservative approach should be
adopted in the time period where accuracy cannot be guaranteed, through publication of
worst-case scenario wave heights or by avoiding any discussion of size for local sources and
advising that people evacuate the largest (most conservative) zone. If wave heights are to
be broadcast, the receiving audience must be educated effectively that wave heights are
given at the shore, and that on-shore inundation height and run-up is likely to be much
greater.

The ideal situation is to avoid any expectation that official warnings will occur in a local
source event at all, thus encouraging total evacuation in response to natural warnings.

GNS Science Report 2012/17 56


2012

3.4.2 Warning systems

Any tsunami warning system in place in Washington or New Zealand that is intended for use
in local warnings must have sufficient redundancy to allow full functionality following a local
earthquake. The experience of effective warning transmission in Minami-Sanriku via
household wireless radio receivers illustrates the effectiveness of a system such as the
NOAA Weather Radio system.

The Japanese approach of utilising siren towers and speakers for post-event
announcements of further events or welfare advice (i.e. locations to receive supplies or
medical treatment) should be considered for post-event response in the United States and
New Zealand where such systems are available. Additional redundancy and resilience is
required in the warning broadcasting system if it is required for further warnings and welfare
announcements. This must be applied to the broadcasting source point, any telemetry, and
the public notification point.

Siren towers and radio-receiver systems are expensive and can be too slow in a local source
event, compared to the natural warning from long or strong earthquakes. These official
warning mechanisms should be treated very cautiously and not expected to be a reliable
solution for local earthquake and tsunami.

3.5 Movements during and after evacuation

3.5.1 Delayed evacuation or non-evacuation

Tsunami education in Washington and New Zealand aims to achieve widespread immediate
evacuation and can benefit from using examples of tsunami tendenko in education
programs, to encourage people to evacuate immediately on experiencing a natural warning.
Whether or not official advice should follow tsunami tendenko in explicitly recommending
that people do not stop to help others is an emotionally and culturally sensitive question and
would require careful consideration before deciding whether or not to implement such advice
in Washington or New Zealand. Assisted evacuation of immobile people or groups requires
careful planning within realistic timeframes for tsunami arrival, and should be regularly
exercised as part of tsunami preparedness activities. There should be clarification of current
FEMA advice (see Section 2.5) to help neighbours evacuate – it is unclear whether this
advice relates to distant tsunami when there would be time to help others without
endangering yourself, or local tsunami when there may not be.

The tsunami tendenko message should be delivered alongside education showing that
delayed evacuation or refusal to evacuate places additional burden on emergency services
or neighbours. The key message is that immediate self-evacuation and prior planning are
the most effective ways to ensure that the maximum number of people survive a tsunami.

3.5.2 Use of motor vehicles during evacuation

Events on March 11th provide support for the approach currently taken in Washington and in
New Zealand, which is to encourage tsunami evacuation on foot wherever possible. Vehicle
congestion was a significant problem in some areas investigated, even where previous
education had encouraged evacuation on foot. Tsunami drills, in which use of vehicles is
banned, appear to have been an effective method of encouraging evacuation on foot and
this should continue to form a key part of education.

GNS Science Report 2012/17 57


2012

Motivating people not to use vehicles is a significant challenge. One strategy in use in
Ryoishi involved ‘stickering’ approved vehicles for the transport of special needs groups,
such as disabled and elderly people. While this is unlikely to be able to be policed in an
event it may well act as a strong education medium (both through selection discussions and
visibility of stickers) and awareness trigger for the need to otherwise avoid cars.

Social responsibility should be harnessed to reduce use of motor vehicles in evacuation by


communicating that use of a motor vehicle puts those in the vehicle and others in danger by
causing congestion and slowing evacuation, and that motor vehicle evacuation should be
reserved for evacuation of immobile residents only.

The emphasis on pedestrian (and bicycle) evacuation in Washington and New Zealand
should be continued, unless a location is a considerable distance from high ground, in which
case, vehicles may be the only way to evacuate coastal zones rapidly enough. In these
cases an efficient scheme is required to streamline large-scale evacuation in vehicles, but
observations from this event have not suggested such a scheme.

3.5.3 Use of boats during evacuation

Further discussion is required to determine the most appropriate recommendation regarding


use of boats to evacuate out to sea. However, Washington and New Zealand should aim to
provide explicit advice to the public on this issue.

It is likely that in many situations following a local tsunami, there would be insufficient time to
launch a boat and reach the open ocean before wave arrival (while likely to be experiencing
very strong currents) and this course of action should not be recommended. Additional
considerations for evacuation by boat are: difficulties returning to port through debris fields or
to damaged moorings, and the ability to stay at sea for many hours.

3.5.4 Schools evacuation

It is vital that the lessons of unnecessary deaths during evacuation of parents and children
are learned and incorporated to current school evacuation advice. The example of children
successfully evacuating Kamaishi City schools provides an excellent example of the benefits
of tsunami education (in this case tsunami tendenko), and this should encourage New
Zealand and Washington to make tsunami education in schools a priority.

Schools should have a school-community approach to tsunami evacuation training, inclusive


of staff, children and parents. All schools in, or in the vicinity of a tsunami hazard zone
should have a tsunami evacuation plan, which is practised regularly with schoolchildren and
teachers, and communicated effectively and regularly to the parents. In this way, all parties
can be educated as to whether a school is at risk of tsunami or not, and what actions should
be taken in the event of a tsunami.

The ideal situation is one in which parents trust that their children are safe while at the
school either because i) they are aware that the school is in a safe location, or ii) they have
confidence that the school staff and their children are all well-trained in evacuation
procedures. This trust could prevent many deaths by removing the requirement or desire for
parents to travel into or through the potential inundation zone when tsunami is imminent. In
the course of reducing the number of parents travelling to schools, potential traffic
congestion and contraflow of traffic can also be reduced.

GNS Science Report 2012/17 58


2012

This school-community approach can also have additional benefits in the education of the
wider community by getting adults involved in disaster education when they might not
otherwise attend community meetings. As one disaster prevention official in T hoku stated,
“The key to effective education is to have everyone envisage the tsunami together and
individuals must take responsibility for their decision of where to go in a tsunami”. He
believes that education in schools in Japan is vital as many children do not attend disaster
workshops in the community, and while there is interest in disaster curriculum in some
schools it is not compulsory; he considers the best way to reach children being through the
curriculum with parents involved in order to create a type of “knowledge permeation
strategy”.

3.5.5 Necessary relocation from evacuation refuges

Tsunami are a dynamic and rapidly evolving hazard, such that a location that at first appears
safe can very quickly become inundated, so people should be empowered to use their
evacuation training and initiative rather than follow a prescriptive evacuation plan to a certain
location and remain there. Personal responsibility is the important factor, and it should be a
key component of education to train people in assessing the changing situation during
evacuation and take action where appropriate.

This education should not come at the expense of ensuring hazard modelling is as accurate
as possible and that any evacuation centres are located outside of the maximum tsunami
inundation extent. Likewise, all vertical evacuation buildings should be designed to remain
safe in the maximum tsunami so that where relocation is not possible, life-safety is still
ensured.

3.5.6 Leaving evacuation refuges too early

Education around the interpretation of arrival times in tsunami warning messages must
continue to stress that arrival times are approximate and that people should not return to
low-lying areas, even if the arrival time has passed without wave arrival. The most important
message is that it may take hours for a tsunami to arrive and the largest wave may be hours
after the first wave.

The importance of receiving an official ‘all-clear’ message must continue to be emphasised


strongly in a consistent fashion across tsunami evacuation literature, particularly in cases
where people may have evacuated due to natural warning and been unable to confirm the
occurrence of a tsunami through official or informal channels.

3.6 Impacts of coastal defence on warning response and evacuation

In Washington and New Zealand it is unlikely that substantial coastal tsunami defences
would be constructed, due to the nature of the coastline and people’s high value on coastal
amenities (e.g. sea views, beach access). The construction of hard defences in response to
a plate boundary local source tsunami hazard should be carefully considered, with the
negative social and environmental impacts, the significant financial investment required, as
well as the estimated levels of physical protection included in any cost-benefit analysis.

GNS Science Report 2012/17 59


2012

Where ‘hard’ defences are considered at even small scales, it should be effectively
communicated to the public that this infrastructure should not be seen as a guarantee of life
safety in any tsunami, and that evacuation remains the appropriate action to take. The issue
of coastal walls blocking the ability to see natural warning signs in the ocean remains.

3.7 Vertical evacuation buildings

The use of vertical evacuation buildings has undoubtedly shown its value in providing safe
refuge during the Great East Japan tsunami. In the locations visited during this research, 37
designated buildings provided refuge to at least 5428 people (see Appendix 4). The use of
buildings is not a replacement for evacuation to high ground, but it does provide effective
alternative options to those unable to evacuate the inundation zone prior to tsunami arrival.
Key lessons can be learnt from the experience of March 11th, for implementation in current
development of vertical evacuation projects in New Zealand and Washington.

3.7.1 Building specification and levels of damage

The specifications for vertical evacuation buildings in Japan, set by Hiroi et al. (2005) were
sufficient in this event with respect to construction type, with damage to designated buildings
in the earthquake and tsunami generally limited to non-structural damage. Along with FEMA
(2008) these guidelines provide a standard for appropriate construction of vertical
evacuation structures, and observations from this event support continuation of the focus on
using reinforced concrete buildings for vertical evacuation. Heavy damage to glazing and
contents should be expected at any inundated facility, while observations in Tōhoku support
the need to find solutions for minimising spread of fire and preventing fire damage.

The appropriate height of buildings is being called into question following instances of waves
almost overtopping buildings. The recommendation of a minimum number of storeys relies
on the underlying hazard modelling for an area, and in order to gain maximum confidence
that a building is tall enough, detailed site-specific tsunami modelling must be carried out at
the site of any proposed facility, using the maximum credible local event. The application of
a factor of safety, such as that applied in FEMA (2008) is still required in addition to the
model-derived estimated maximum tsunami height to account for uncertainty in such
estimates. A blanket requirement for a single minimum height may not be the right approach,
given the varying flow depth with onshore topography seen in Japan.

3.7.2 Community engagement

Community engagement on the selection of vertical evacuation buildings should be


encouraged. It is vital that members of the local community are familiar with the buildings as
part of the community preparedness plan, and are aware that the buildings are proposed as
a secondary option to evacuation to high ground. If a project has the support of the
community it will likely increase co-operation with building owners who may initially be
reticent about the designation of their building.

Designated facilities should be incorporated into tsunami evacuation drills to foster familiarity
with their use. Care should be taken to continue to prioritise evacuation to high ground, but
evacuation to vertical evacuation facilities should be practised where a local tsunami is
known to be likely to prevent people in certain locations reaching high ground.

GNS Science Report 2012/17 60


2012

3.7.3 Evacuation building signage

Consistent signage should be adopted and applied to all buildings designated for vertical
evacuation use. These signs should be clearly displayed at the top of buildings and above
entrances, indicating the most appropriate access route to upper stories. The signage should
be consistent with current styles and messages in use in the local area. The signage is
intended to allow easy identification for local people and visitors to the area (or emergency
responders), and enhances awareness of the facilities.

3.7.4 Owner agreement

Owner agreement is likely to be a significant issue in New Zealand and Washington if private
buildings are considered for use. Owner agreement issues generally focus on responsibility
or liability for evacuees and issues of access. Leveraging community interest and
encouraging owners to see the provision of the facility as a benefit to the community has
been shown in Japan to be an effective way to gain agreement of building owners.

3.7.5 24-hour access

The ability to quickly access vertical evacuation buildings at any time of day or night is vital.
The community should be aware that facilities are available at night, and building owners
should be prepared to allow access to the building at any time. Direct external access from
the ground floor to upper storeys (above 3rd or 4th storey) or the roof should be provided
where possible, but where this is unfeasible, agreement should be sought between the
building owners and the community in recognising that in an emergency, people may need
to gain access by other means.

Promoting the use of “key-holders” – representatives who have keys with which to access
designated buildings in the case of an emergency when the building is locked appears to be
an effective method, but requires responsibility on the part of key-holders to go to the
building in the event of a tsunami. These people would ideally be civil defence volunteers or
people in a similar role. Allowing evacuees to enter by force should be investigated as a
possibility, but is likely to be an inappropriate solution for buildings containing sensitive data
such as public offices or commercial premises. In each case, dialogue between building
owners and the community is encouraged in order to find the most appropriate solution.

3.7.6 Evacuee welfare

In any vertical evacuation building, there should be provision of shelter, food and water, and
communication links to civil defence and emergency services at upper storeys. Provision of
dedicated supplies has been shown to be extremely important given the potential for refuges
to be in use for up to a few days if tsunami waters subside slowly or debris traps evacuees.
Adequate provisions may help reduce any need or urge of evacuees to leave the refuge
earlier than necessary. The amount of provisions should be commensurate with an
estimated number of occupants, which should be researched through a combination of
evacuation mapping and travel time mapping to estimate the likely level of occupancy.
However, these provisions should be managed in a way so as not to not lessen the need for
the public to prepare personal emergency kits with their own supplies.

GNS Science Report 2012/17 61


2012

3.8 Additional recommendations from observations and interviews

It is important to provide education on tsunami risk and response to people living inland, as
well as those living at the coast. Everyone should have the awareness to be able to take
responsibility for their own actions when in a tsunami-risk area. If the awareness exists in
visitors, it is likely they will be able to respond to any warnings and signage more effectively,
even in an unfamiliar area.

Landslide hazard should be taken into account for tsunami evacuation, in areas where high
ground is reached by steep inclines or up narrow valleys. This is important for the
occurrence of local tsunami with significant prior ground shaking, which could trigger
landslides which block evacuation routes. The awareness to respond and use an alternative
evacuation route could save many lives.

The response of some volunteer fire-fighters in Japan was to wait to confirm the occurrence
of tsunami arrival on television before responding to the emergency. The previous repeated
closure of tsunami gates with no subsequent event may have contributed to some
complacency among fire-fighters in T hoku. It is unclear whether this slowed the fire
services response to closing tsunami gates in this event, but this effect must be mitigated to
ensure the most rapid response possible following every warning or significant local
earthquake. Training of emergency personnel must emphasise the need for immediate
response to natural warnings.

4.0 ACKNOWLEDGEMENTS

This research was commissioned by the Washington Military Department Emergency


Management Division and supported by New Zealand Ministry of Science and Innovation
core geohazards funding to GNS Science, and was conducted in collaboration with
Japanese colleagues.

The authors would like to acknowledge the officials and researchers who were interviewed
during the field survey for their time, enthusiasm and detailed consideration of our research
questions, and Joel Challender for translation.

5.0 REFERENCES

Bombaadi. (2011). Rare Footage of Tsunami. YouTube. Retrieved November 21, 2011, from
http://www.youtube.com/watch?v=gzktUwyYWNM

Cabinet Office Government of Japan. (2007). Disaster Management in Japan (p. 46). Tokyo.
Retrieved from http://www.bousai.go.jp/1info/pdf/saigaipanf_e.pdf

Central Disaster Prevention Council. (2008). Earthquake in Japan Trench Quantitative


Mitigation strategy. Tokyo: Cabinet Office Government of Japan. Retrieved from
http://www.bousai.go.jp/jishin/chubou/taisaku_kaikou/pdf/081212_01.pdf

City of Sendai. (2010). Chile Earthquake Tsunami Evacuation Survey Summary (p. 16).
Sendai. Retrieved from
http://www.city.sendai.jp/kurashi/shobo/shiryo/1196975_1390.html

GNS Science Report 2012/17 62


2012

Clallam County Emergency Management. (n.d.). Evacuation routes for Clallam Bay and
vicinity. Clallam County: Clallam County. Retrieved from
http://www.clallam.net/Maps/documents/ClallamBaymap.pdf

Disaster Mitigation Section funato City. (n.d.). Outline of Earthquake and Tsunami Disaster
Countermeasures in funato City. funato City, Iwate Prefecture.

Dowrick, D. J. (1996). The Modified Mercalli Intensity Scale - Revisions Arising From Recent
Studies of New Zealand Earthquakes. Bulletin of the New Zealand Society for Earthquake
Engineering, 29(2), 92-106.

Earthquake Engineering Research Institute (EERI). (2011a). The March 11, 2011, Great
East Japan (Tohoku) Earthquake and Tsunami: Societal Dimensions.

EEFIT. (2011). The Mw9.0 T hoku earthquake and tsunami of 11th March 2011 - A field
report by EEFIT. London, UK.

Expert Committee on the Working Group on Tsunami disaster evacuation. (2010). Summary
of major tsunami. Cabinet Office (Disaster Management Division) Government of Japan.
Retrieved from http://www.bousai.go.jp/jishin/chubou/taisaku_tsunami/1/index.html

FEMA. (2008). Guidelines for Design of Structures for Vertical Evacuation from Tsunamis,
FEMA Report 646. Redwood City, California.

FEMA. (2011). Tsunamis. Retrieved December 1, 2011, from http://www.ready.gov/tsunamis

Fire and Disaster Management Agency of Japan. (2012). Report 144 on the effects of the
March 11, 2011 Great T hoku earthquake, February 14 2012. Retrieved February 16,
2012, from http://www.fdma.go.jp/bn/2012/detail/691.html

Geospatial Information Authority of Japan (GSI). (2011). The 2011 off the Pacific Coast of
T hoku Earthquake: Crustal Deformation and Fault Model. Retrieved November 18,
2011, from http://www.gsi.go.jp/cais/topic110422-index-e.html

GeoSpatial Information Authority of Japan. (2011b). Year 2011 (2011) due to earthquake
subsidence investigation northeastern Pacific Ocean off the coast. Retrieved July 20,
2011, from http://www.gsi.go.jp/sokuchikijun/sokuchikijun40003.html

Harding, D., and Harding, B. (2011). Tendenko: Surviving the Tsunami. Al Jazeera.
Retrieved from
http://www.aljazeera.com/programmes/witness/2011/11/20111114121620284203.html

Hawke’s Bay CDEM Group. (2011, September). Latest hazard maps inform tsunami
preparedness. Impact, 8-10. Napier, New Zealand.

Hiroi, O., Abe, K., Kabeyasawa, T., Kawata, Y., Sugano, T., Tanaka, A., Yoshimura, H., et
al. (2005). Guideline for Tsunami Evacuation Buildings. Tokyo. Tsunami Evacuation
Building Guideline Committee, Government of Japan Cabinet Office. Retrieved from
http://www.bousai.go.jp/oshirase/h17/tsunami_hinan.html

Imamura, F. and Abe, I. (2009). History and Challenge of Tsunami Warning Systems in
Japan. Journal of Disaster Research, 4(4), 267-271.

GNS Science Report 2012/17 63


2012

International Tsunami Information Center. (2006). What to Do? Tsunami Safety for Boaters.
Honolulu, Hawaii: UNESCO IOC ITIC. Retrieved from
http://ioc3.unesco.org/itic/files.php?action=viewfile&fcat_id=270&fid=411

Ishinomaki City. (n.d.). Ishinomaki District (Tsunami Hazard Map). Retrieved December 16,
2011, from
http://www.city.ishinomaki.lg.jp/static/GUIDE_ishinomaki/tsunami/tsunami-z.html

Japanese Meteorological Agency. (2006). Tsunami Warning /Advisory and Tsunami


Information. Retrieved November 25, 2011, from
http://www.seisvol.kishou.go.jp/eq/eng/fig/tsunamiinfo.html

Japanese Meteorological Agency. (2011a). The 2011 off the Pacific coast of T hoku
Earthquake -Portal-. Retrieved November 18, 2011, from
http://www.jma.go.jp/jma/en/2011_Earthquake.html

Japanese Meteorological Agency. (2011b). Tables Explaining the JMA Seismic Intensity
Scale. Retrieved November 17, 2011, from
http://www.jma.go.jp/jma/en/Activities/inttable.html

Japanese Meteorological Agency. (2011c). Seismic Intensity Database - search facility.


Retrieved November 18, 2011, from
http://www.seisvol.kishou.go.jp/cgi-bin/shindo_db.cgi?pref=0&START_DATE=

Japanese Meteorological Agency. (2011d). Monitoring of Earthquakes, Tsunamis and


Volcanic Activity. Retrieved November 25, 2011, from
http://www.jma.go.jp/jma/en/Activities/earthquake.html

Japanese Meteorological Agency. (2011e). JMA monthly report for earthquakes and
volcanoes, March 2011 (pp. 58-63). Tokyo. Retrieved from
http://www.seisvol.kishou.go.jp/eq/gaikyo/monthly201103/20110311_T hoku_1.pdf

Japanese Meteorological Agency. (2011f). The 2011 off the Pacific coast of T hoku
Earthquake - Location of the Main Shock and Aftershocks (p. 1). Retrieved from
http://www.jma.go.jp/jma/en/2011_Earthquake/2011_Earthquake_Aftershocks.pdf

Johnson, V. A. (2011). Disaster Preparedness Education in Schools : Recommendations for


New Zealand and the United States. Public Policy. Wellington, New Zealand.

Johnston, D., Paton, D., Crawford, G. L., Ronan, K., Houghton, B., and Borgelt, P. (2005).
Measuring Tsunami Preparedness in Coastal Washington, United States. Natural
Hazards, 35(1), 173-184. doi:10.1007/s11069-004-2419-8

Johnston, D., Becker, J., Gregg, C., Houghton, B., Paton, D., Leonard, G. S., and Garside,
R. (2007). Developing warning and disaster response capacity in the tourism sector in
coastal Washington, USA. Disaster Prevention and Management, 16(2), 210-216.
doi:10.1108/09653560710739531

Johnston, D., Becker, J., Gowan, M., Leonard, G. S., and Saunders, W. (2009). Evaluating
warning and disaster response capacity in the tourism sector in Long Beach and Ocean
Shores, Washington, USA.

Kunugi, T. (2000). Relationship between Japan Meteorological Agency instrumental intensity


and instrumental Modified Mercalli intensity obtained from K-NET strong-motion data.
Zisin, 53, 89-93.

GNS Science Report 2012/17 64


2012

Kurowski, M. J., Hedley, N., and Clague, J. J. (2011). An assessment of educational tsunami
evacuation map designs in Washington and Oregon. Natural Hazards, 59(2), 1205-1223.
doi:10.1007/s11069-011-9780-1

Leonard, G. S., Johnston, D. M., Gregg, C., Garside, R., Saunders, W. S.A., Becker, J. and
Fraser, S. (2011). Evaluation and recommendations for tsunami and lahar exercises in
Washington State., GNS Science Report 2011/32, 51 p.

MCDEM. (2008a). Tsunami Evacuation Zones. Director’s Guideline for Civil Defence
Emergency Management Groups [DGL 08/08]. (p. 19). Wellington, New Zealand.

MCDEM. (2008b). National Tsunami Signage Technical Standard for the CDEM Sector [TS
01/08] (pp. 1-20). Wellington, New Zealand.

MCDEM. (2010a). Tsunami Advisory and Warning Plan. Supporting Plan [SP01/09] Revised
October 2010. (p. 64). Wellington, New Zealand.

MCDEM. (2010b). Working from the same page: Consistent Messages for CDEM. August
2010. Wellington, New Zealand.

MSN Sankei News. (2011, April 14). Miracles of Kamaishi as a result of following “Three
principles of evacuation”:Students all safe thanks to disaster reduction education. MSN
Sankei News, Translated by SEEDS Asia. Retrieved from
http://sankei.jp.msn.com/life/news/110413/edc11041314070001-n1.htm

Minoura, K., Imamura, F., Sugawara, D., Kono, Y., and Iwashita, T. (2002). The 869 Jogan
tsunami deposit and recurrence interval of large-scale tsunami on the Pacific coast of
northeast Japan. Journal of Natural Disaster Science, 23(2), 83-88.

Miyagi Prefectural Government. (2004). Tsunami Hazard Map (Minamisanriku). Retrieved


November 25, 2011, from http://www.pref.miyagi.jp/sabomizusi/bousai/bou-ht5-1.html

Murakami, H., and Kashiwabara, K. (2011). Travel Means for Tsunami Evacuation in the
2011 T hoku Pacific Ocean Earthquake - Questionnaire Survey in Natori City. Annual
Conference of Institute of Social Safety Science (p. 4).

NOAA/NWS. (2011). TsunamiReadyTM Helps Communities at Risk. Retrieved December


15, 2011, from http://www.tsunamiready.noaa.gov/

NPO CeMI. (2011). Northeastern Pacific Ocean off the coast earthquake: Preliminary
analysis of questionnaire survey on tsunami (p. 22). Tokyo, Japan.

NTHMP Mitigation and Education (MES) Subcommittee. (2011). Guidelines and Best
Practices for Tsunami Evacuation Mapping Guidelines.

National Police Agency of Japan. (2012). Damage Situation and Police Countermeasures
associated with 2011 T hoku district: 31st January 2012. Retrieved February 1st 2012,
from http://www.npa.go.jp/archive/keibi/biki/higaijokyo_e.pdf

Ohtake, M., and Ueda, H. (2002). Statistical Forecast of the Next Miyagi- Oki , Northeast
Honshu Earthquake. T hoku Geophysical Journal (Sci. Rep. T hoku University Series 5),
36(3), 299-310. Retrieved from
http://ir.library.T hoku.ac.jp/re/bitstream/10097/45406/1/AA00459426021057.pdf

GNS Science Report 2012/17 65


2012

Ozaki, T. (2011). Outline of the 2011 off the Pacific coast of Tohoku Earthquake (Mw 9.0) -
—Tsunami warnings/advisories and observations. Earth Planets Space, 63(7), 827-830.
doi:10.5047/eps.2011.06.029

Pacific County Emergency Management. (2007). Tsunami! Evacuation Map for Long Beach
and Ilwaco. Pacific County. Retrieved from
http://www.co.pacific.wa.us/pcema/Documents/LongBeachTsunami.pdf

Project Safe Haven. (2011a). Project Safe Haven: Tsunami vertical evacuation on the
Washington coast - Grays Harbor County (p. 93). Seattle, Washington.

Project Safe Haven. (2011b). Project Safe Haven: Tsunami vertical evacuation on the
Washington coast - Pacific County (p. 89). Seattle, Washington.

Sagara, J. (2011). Critical Cause Analysis of Delayed Evacuation in the Great East Japan
Earthquake and Tsunami. Presentation at IRDR International Conference, Beijing. 1
November, 2011.

Sato, M., Ishikawa, T., Ujihara, N., Yoshida, S., Fujita, M., Mochizuki, M., and Asada, A.
(2011). Displacement above the hypocenter of the 2011 T hoku-Oki earthquake. Science
(New York, N.Y.), 332(6036), 1395. doi:10.1126/science.1207401

Saunders, W.S.A.; Prasetya, G.; Leonard, G.S. 2011 New Zealand's next top model :
integrating tsunami inundation modelling into land use planning. Lower Hutt: GNS
Science. GNS Science miscellaneous series 34. 37 p.

Statistics Bureau. (2010). 2005 Population Census: Population of Japan (Final Report of the
2005 Population Census). Basic Figures for Shi, Ku, Machi and Mura (Table 82).
Retrieved from http://www.e-stat.go.jp/SG1/estat/ListE.do?bid=000001025191&cycode=0

Suganuma, K. (2006). Recent Trends in Earthquake Disaster Management in Japan.


Science and Technology Trends, 19.

Takahashi, S., Kuriyama, Y., Tomita, T., Kawai, Y., Arikawa, T., Tatsumi, D., and Negi, T.
(2011). Urgent Survey for 2011 Great East Japan Earthquake and Tsunami Disaster in
Ports and Coasts - Part I (Tsunami) - An English Abstract of the Technical Note of Port
and Air Port Research Institute, No. 1231, April 28. (p. 9).

Tar Town Municipal Government. (n.d.). Tar Town Tsunami Hazard Map. Tar Town.

Unknown. (2011a, April). Tsunami hit more than 100 designated evacuation sites. The
Japan Times. Kyodo. Retrieved from http://www.japantimes.co.jp/text/nn20110414a4.html

Unknown. (2011b, May 30). 10,000 people survive after evacuating to a (designated
evacuation) building. Iwate Nichi Nichi Shinbun, 7. Ichinoseki, Japan.

Venturato, A. J., Arcas, D., and Kanoglu, U. (2007). Modeling Tsunami Inundation From a
Cascadia Subduction Zone Earthquake for Long Beach and Ocean Shores, Washington.
(p. 26). Seattle, WA.

Washington Military Department Emergency Management Division. (1999). Tsunami! Safety


Tips for the Washington Coast! Evacuation Map for Grays Harbor and Pacific Counties.
Camp Murray, Washington: Washington Military Department Emergency Management
Division. Retrieved from
http://www.emd.wa.gov/publications/pubed/What_is_a_tsunami.pdf

GNS Science Report 2012/17 66


2012

Washington State Department of Natural Resources. (2007). Tsunami! Evacuation Map for
Ocean Shores and Vicinity. Washington, United States: Washington State Department of
Natural Resources. Retrieved from
http://www.dnr.wa.gov/Publications/ger_tsunami_evac_oceanshores.pdf

West Coast and Alaska Tsunami Warning Center. (2011). Tsunami Safety Rules. Retrieved
December 19, 2011, from http://wcatwc.arh.noaa.gov/about/safety.php

GNS Science Report 2012/17 67


2012

APPENDICES

APPENDIX 1 INTERVIEW QUESTIONS (ORIGINAL ENGLISH VERSION)

Research questions to be used in interviews with disaster management officials in T hoku,


Japan.

The aim is to get an understanding of the role of natural and informal warnings, and vertical evacuation buildings
th
in the tsunami mitigation strategy of several municipalities, and to understand their performance in the March 11
2011 T hoku tsunami. Lessons learned will inform further work on tsunami evacuation in Washington State, U.S.,
and New Zealand as part of an on-going Ph.D. project and collaboration between GNS Science and Washington
State Emergency Management Department.

Preparedness – Vertical Evacuation (‘V.E.’)

1. Was V.E. available in the city prior to the 2011 tsunami?


a. Was this in response to national or prefectural government directive – as part of the Basic Disaster
Plan, or municipality initiative?

2. How many buildings were designated for use as a V.E. facility?


a. When were the buildings designated?
b. How were buildings chosen? (basis of height and material, or detailed analysis by engineers?)
c. What design tsunami was for used basis of hazard in the designation process? modelled Miyagi-
ken-oki, Chile 1960, Meiji Sanriku 1933?
d. Has there been resistance from building owners when designating buildings as V.E. buildings?
e. Were additional modifications made to any of these buildings when designated? (retrofit of
structure, fitting of railings, external stairs, sirens) Is this a condition of them being designated for
vertical evacuation?
f. Are there other styles of vertical evacuation facilities? (Artificial berms, towers)

3. What style of evacuation map is available for the city? (see example and obtain copy where available)
a. When were the evacuation maps developed?
b. What was the design tsunami? Miyagi-ken-oki, Chile 1960, Meiji Sanriku 1933?
c. Were these developed with V.E. in place, or was development separate?
d. Are V.E. facilities explicitly marked on the evacuation maps? (e.g. Ishinomaki has a single symbol,
Kesennuma are different)

4. How are the public made aware of V.E. facilities?


a. What is the educational advice given to the public? (“high ground first, then V.E. building…”?)
b. Is V.E. included in tsunami literature provided by management office?
c. Do evacuation drills involve the use of V.E. buildings?

Preparedness – Natural/Informal Warnings

5. How are natural warnings signs presented in tsunami education and mitigation?
a. Are the public aware of the importance of natural warnings signs?
b. Does literature use an event of certain duration/level of ground shaking as a threshold for when
people should evacuate?

6. Are the public made aware of the possibility of informal warnings in the event? (warnings from neighbours…)

On 11 March 2011 – Warnings and evacuation response

7. How long did ground shaking last in the town?


a. Does this satisfy advice given for reacting to natural warnings?

8. What was the JMA intensity in the town, or personal experience of ground shaking?
a. Does this satisfy advice given for reacting to natural warnings?

9. What was the arrival time of first wave in the town?


a. How large was this first wave?
b. Arrival times of larger waves?

10. What time did the official warning commence, and how was this communicated?

GNS Science Report 2012/17 68


2012

a. How did the disaster management initiate the warning? (official message from JMA/government, or
as a result of natural/informal warnings?)
b. Were there power outage issues, any other issues that prevented communication of the warning?

11. When did people begin to evacuate?


a. Was there a large response to ground shaking or did people wait until wave arrival?
b. At what point do people report beginning evacuation – ground shaking commencing, ground
shaking ceasing or tsunami wave arrival?
c. What proportion of the population ignored natural warning signs and waited for the tsunami
warning, or did not evacuate?

12. Where did the population evacuate to? (get proportions and age distribution if possible)
a. How many people evacuated to designated high ground?
b. Other high ground?
c. Vertical evacuation buildings?
d. Upper floors of other buildings?

13. How did people evacuate? (car, on foot, bicycle).

14. How far did people travel to each evacuation place?


a. Were evacuation times/distances consistent with estimations of average walking times?

15. Where were issues of congestion observed, and did this result in further casualties (i.e. were many people
killed in their cars?)

On 11 March 2011 – Vertical Evacuation (V.E.)

16. How many designated V.E. structures were available in the town in March 2011?
a. Which had signage indicating their emergency use?
b. Were all of these buildings used?
c. Were additional (non-designated buildings) used by large groups of people?
d. What is the nature of arrangements with buildings owners, and who is liable for safety of evacuees
in the building?
e. Are there official personnel at each building in the event, to direct/help evacuees?

17. During tsunami warnings, were the public specifically advised to evacuate to any vertical evacuation
building? (or to “upper floors of a tall building”)

18. How many people evacuated into each of the vertical evacuation buildings (designated/non-designated)?
a. How long did people remain in the buildings?
b. What distances did they travel to each building?
c. Do any of the buildings have provisions for medium-term occupancy? (timeframe: days; provisions:
food and water, radio communication, shelter)

19. Issues with the evacuation buildings


a. Was there any overcrowding reported in a building?
b. Were there any buildings that were inaccessible prior to inundation? (due to being locked, or
earthquake-damaged)
c. How many of the buildings are estimated to be locked and inaccessible at night?
d. Did any buildings suffer damage/flooding that caused further evacuation? (i.e. fire damage, debris
impact)
e. Did any buildings suffer damage/flooding that caused deaths of occupants?
f. Did people refrain from using such buildings following such a large earthquake? (what is public
perception of safety of such buildings post-earthquake)

Since March 2011 – Vertical Evacuation (V.E.)

20. Has there been a review of vertical evacuation facilities since March 2011?
a. Is this municipality level review, or national, prefectural?
b. Is there any national/uniform guidance from national government on application of vertical
evacuation?

21. Are there plans in place to change the (vertical) evacuation strategy for the town? (new structures being
built, re-assigning building use, more emphasis on vertical evacuation in the evacuation strategy)

GNS Science Report 2012/17 69


2012

APPENDIX 2 INTERVIEW QUESTIONS (JAPANESE TRANSLATION)

Research questions to be used in interviews with disaster management officials in T hoku,


Japan. These questions were translated and provided to interviewees in preparation for the
interviews.

垂直避 耐震基準を満 ニャ 避 /略V.E. 準備 い


. 津波以前 、こ 垂直避 を るこ 能 ?取
ン 能 ら、 れ 基本災害計 一環 、国家 県 政府
よる指示 よる ? 自没体独自 よる ?取

.垂直避 用 指定され ニャ 何件あり ?取


ン取 よう ニャ 指定され い ?取
ン取垂直避 用ニャ よう 選 れ ?ニャ 高さ、使用され いる
建築資材、 技術者 よる詳細 分析 よる ?取
ン取垂直避 用 ニャを指定 る 、 よう 災害を考慮され ?想
定宮城県沖津波、 チモ地震、 明没 陸津波
ょう ?取
ン取取
垂直避 用ニャ 指定 る 、ニャ 所暼者 ら反対を けるこ あり
?取
ン取垂直避 用 ニャ 指定され 際 、ニャ 変更を加えるこ あり
? 構造 改築、柵を立 る、外側 階段を作る、サイヤンを ける

ン取ニャ 変更を加えられるこ 、垂直避 用ニャ 指定 る条件 ?取
ン取 他、垂直避 用ニャ よう 特徴 あり ? 人口 小段や
タワヴ 取

.こ よう 避 用地図 使用され い ? サンプャ あれ


せ 、 能 あれ コヌヴを入手 取
ン取避 用地図 い ?取
ン取避 用地図を設計 る際 、 津波 考慮され ?想定宮城県沖津
波、 チモ地震、 明没 陸津波 ょう
?取
ン取避 用地図 、垂直避 用ニャ 指定 伴 作られ ?
別々 設計され ?取
ン取避 用地図 、垂直避 用ニャ 明確 表示され い ? 石巻
統一され クンボャ 使用され い 、気 沢 使用され い

.垂直避 用ニャ 関 、人々 よう 知らされ いる ょう ?取


ン取 よう 避 関 る指針 あり ? 高 避 、 後垂
直避 用ニャ 移動 る 取 取

GNS Science Report 2012/17 70


2012

ン取災害対策本部 よる津波 関 る文献 中 、避 用ニャ い 記載され


い ?取
ン取避 訓練 際 、垂直避 用ニャ 避 行い ?取
自然 ら 警告、非公式 警告 関 る準備 い 取

.津波教育、 災害被害を軽減 る 、自然 ら 警告 い 、 よ
う 表現され い ?取
ン取人々 よう 自然 ら 警告を深刻 け い ?取
ン取津波 関 る文献 中 、地震 ある一定 長さや震度 達 れ 避
書 れ い ?取

.こ よう 地震ン津波 際 、近所 人 ら 警告 、非公式 形 警告
人々 注意を払 い ?取

暻 日 ら取 警告及び避 行動 い 取

.こ 、揺れ らい続 ?取
ン取こ 長さ 、避 自然 ら 警告 感 ?取

.こ 気象庁 観測 震度 い ? 、個人的 ら
い 強さ 感 られ ?取
ン取こ 地震 強度 、避 自然 ら 警告 感 ?取

.こ を最初 津波 襲 何時 ?取
ン取最初 津波 規模 らい ?取
ン取より大 い津波 到達 何時 ?取

. 式 津波警報 出され 何時 ? 警報 よう 形 伝え
られ ?取
ン取災害対策本部 、 よう 警報を出 ? 気象庁 よる通達
を け 、政府 よる通達を け 、 自然及び非公式 警告を け
出 ? 取
ン取停電 よる問題、 警報を阻 るよう 他 問題 あり
?取

.い 、人々 避 を始 ?取
ン取地震 後 大勢 避 を始 ? 最初 津波 到達 る 待
い ?取
ン取 時点 避 を始 報告され い ?地震 起 、揺れ
さ 、 津波 到達 ょう ?取
.自然 ら 警告を無視 津波 到着 る 待機 人、 避
人 人口 何割 らい ?取

GNS Science Report 2012/17 71


2012

.人々 こ 避 ? 割合、 代 分布 分 る
あれ 入手 取
指定され 高 何人 らい 人 避 ?取
ン取 他 高 避 人 い ?取
ン取 垂直避 用ニャ 避 人 い ?取
ン取垂直避 用ニャ以外 建物 階 避 人 い ?取

.人々 よう 避 ? 車、徒歩、自転車 取

.避 場所 らい 距 を移動 ?取
ン取避 時間 避 距 、 均的 歩 早さ 矛盾 あり せん ?
アメモィ合衆国連邦緊急事態管理局 日本 取

.渋滞ン混雑 問題 よう 場所 見られ ?こ 問題 よ 、
死傷者数 増え ? 例え 、大勢 人 車 中 亡 り ?


暻 日 ら取 垂直避 用ニャ い 取

. 暻 時点 、こ 垂直避 用ニャ 何件あり ?取
ン取緊急時 避 る建物 あるこ 表示され いるニャ あり ?取
ン取これら全 ニャ 避 使われ ?取
ン取避 用 指定され い いニャ 、大勢 人々 避 ?取
ン取ニャ 所暼者 よう り決 あり ?ニャ 避 者 安全
、誰 責任を負い ?取
ン取 れ れ ニャ 、避 者 指示を出 り助け り る、公 責任者
い ?取

.津波警報 際 、特 垂直避 用ニャ 高層ニャ 層階 避 るよ
う、避 指示 出され ?取

.垂直避 用ニャ 避 何人 ? 垂直避 用ニャ、及び れ
以外 ニャ れ れ何人 ? 取
ン取 れ らい 期間、人々 ニャ い ?取
ン取ニャ 移動距 らいあり ?取
ン取避 者 数日 期間滞在 る場合、食料や水、メグアや避 所 、支給物
資等を提供 るニャ あり ?取

.避 用ニャ 問題 い 取
ン取混雑 いる 報告され ニャ あり ?取

ン取浸水前 、鍵 い 、地震 壊れ い 理由 、入れ いニ


ャ あり ?取

GNS Science Report 2012/17 72


2012

ン取夜間 鍵 い る 入るこ いニャ 何件 らいあ


る 見積 られ い ?取

ン取火事や瓦礫 よる損傷 地震や津波 よる被害を け、避 人々


さら 避 けれ ら い状況 ニャ あり
?取

ン取地震や津波 よる被害を け、避 者 亡 ニャ あり


?取

ン取あれほ 大地震 後、人々 こ よう ニャ 入る を らい


? 地震後、こ よう ニャ 安全性 関 る意識 よう 変わり
? 取


暻 ら取 垂直避 V.沼. い 取

. 暻 ら、垂直避 用 施設 い 再調査され ?取
ン取 再調査 、自没体 よる ? 県、国 よる ?取

ン取垂直避 用 施設を設定 る際 、日本政府 よる国ヤベャ 統一され 指


針 あり ?取

.こ い 、 垂直 避 戦略を変更 る予定 あり ? 新
い構造を設計 いる、ニャ 使用 関 再割当を いる、避 戦略
い 垂直避 をより重視 るよう 等 取

GNS Science Report 2012/17 73


2012

APPENDIX 3 SUMMARY OF KEY POINTS MENTIONED IN INTERVIEWS

The following tables provide a summary of points that were raised in each location during
our interviews. Due to the nature of semi-structured interviewing, the absence of a point
being raised does not necessarily mean that it does not apply in that location, rather, this
point did not emerge from that particular interview. It is common for topics to emerge in early
interviews, which then appear in later interviews. The order of locations in this table reflects
the order of interviews during our investigation; our first interview took place in Tar Town
and the final interview in Natori City.

Kesennuma City

Minami-Sanriku

Ishinomaki City
Kamaishi City
Tar Town

Natori City
funato City

Town
Earthquake and tsunami

Recorded JMA seismic intensity 5- to 5+ to


6- 6- 6- 6- 6+
(Japanese Meteorological Agency 2011c) 5+ 6-

Earthquake felt more than two minutes long


*Reported as longer than 1 minute of shaking • • • • • * •
Earthquake felt unusual or strongest ever *a question
asking what the earthquake felt like emerged after the
first few interviews
• • • • •

Expectation among officials that it was Miyagi-ken-oki • • •


Reported public did not expect large tsunami, despite
strength • • • •
Drawdown observed *this was not an interview question,
but was offered as an observation by the interviewee •
Stated first tsunami waves arrived in under 30 mins • • • •
Mentioned largest waves were not first • • • • •
Kesennuma City

Minami-Sanriku

Ishinomaki City
Kamaishi City
Tar Town

Natori City
funato City

Town

Preparedness

1896 1896 1896 1896


1896
Significant damage to city/town in previous events 1933 1933 1933 1933
1933
1960 1960 1960 1960

Interviewee felt there was high awareness that EQ means


tsunami • • •
Stated previous events/evacuations with small or no
tsunami –leading to complacency and normalisation • • • •

GNS Science Report 2012/17 74


2012

Kesennuma City

Minami-Sanriku

Ishinomaki City
Kamaishi City
Tar Town

Natori City
funato City

Town
Earthquake and tsunami

Interviewee pointed out that local and distant source


events are differentiated in preparedness activities • •
Reported education encourages response to natural
warnings • • • •
Mentioned local volunteer disaster groups operate • • •
Stated drills held, additional to annual national tsunami
drill • • •
Reported community involved in making evacuation maps • • • •

Kesennuma City

Minami-Sanriku

Ishinomaki City
Kamaishi City
Tar Town

Natori City
funato City

Town
Tsunami Warning

Reported loss of power for warning system


* due to inundation, rather than earthquake • • •* •* •
Mentioned estimated wave height included in warning • •
Stated estimated wave height specifically not included in
audible warning • •
Stated informal warnings (e.g. one person mentioned to
another) saved people
* including seeing others running
•* •* • • •
Kesennuma City

Minami-Sanriku

Ishinomaki City
Kamaishi City
Tar Town

Natori City
funato City

Town

Evacuation

Reported significant numbers evacuated due to ground


shaking • • • • •
Stated significant road congestion due to vehicle use • • • •
Mentioned people returned home before warning
cancellation • •
Reported parents collected children from school resulting
in deaths • • • •

GNS Science Report 2012/17 75


2012

Kesennuma City

Minami-Sanriku

Ishinomaki City
Kamaishi City
Tar Town

Natori City
funato City

Town
Earthquake and tsunami

Mentioned people waited for family members before


evacuating •
Reported deaths of citizens checking on others • • •
Discussed deaths of people with responsibilities - carer,
firemen • • •
Reported people close to the inland hazard boundary
evacuated late or didn’t evacuate • • •
High ground exists within 1km of inundated area (approx.
15 minutes walking time for an able person – a
conservative approach to values from FEMA 2008)
• • • • • •
Reported relocation of evacuees after reaching first
refuge • • •
Mentioned people returned to low-lying areas when no
wave arrived by estimated arrival time • •
Stated defences caused complacency in response
*Converse applied – low walls did not give impression of
safety
• • *

Mentioned defences obstruct view of water • •


Kesennuma City

Minami-Sanriku

Ishinomaki City
Kamaishi City
Tar Town

Natori City
funato City

Town

Vertical Evacuation

Stated existing buildings officially designated • • • • •


Mentioned new buildings constructed with V.E. planned • •
Noted inconsistent signage of buildings • • • •
Noted that community suggested additional bldgs.
*officially designated, ^ not officially designated •^ •^ •* •*
Mentioned community-recommended buildings were
successfully used in vertical evacuations • • •
Reported vertical evacuation saved lives • • • • • • •
Recognition that water/food stocks required • • • •

GNS Science Report 2012/17 76


2012

Kesennuma City

Minami-Sanriku

Ishinomaki City
Kamaishi City
Tar Town

Natori City
funato City

Town
Earthquake and tsunami

Reported that required height for vertical evacuation


should be >5-storeys, due to inundation close to roof level
of some vertical evacuation buildings
• •

Mentioned people remained in buildings for >24 hours • • •


Reported 24-hr access by building function/structural
feature • • • • •
Noted 24-hr access by community arrangement • • • •
Reported overtopping of an officially designated vertical
evacuation building overtopped •
Recognition that emergency communications links to civil
protection are required in vertical evacuation buildings • •

GNS Science Report 2012/17 77


2012

APPENDIX 4 SUMMARY OF VERTICAL EVACUATION BUILDINGS IN LOCATIONS VISITED

Inundated
City/town Building Construction Height People saved Signage 24-hour access
storeys

Minami- Sanriku Shizugawa Hospital RC 5-storey 4 320 No Open 24-hrs

Matsubara community
Minami-Sanriku RC 4-storey 4 44 Yes Residents open
apartment block

Minami- Sanriku Takano Kaikan RC 4-storey 3 330 No External stairs

Minami- Sanriku Fishing cooperative RC 2-storey 2 - No Unknown

Ishinomaki York Benemaru RC 1-storey 1 No External stairs

External stairs to
Ishinomaki Homac Steel frame 1-storey 1 500 No
roof car park

Ishinomaki Hotaru funeral centre Unknown 2-storey 1 No Unknown

2 private commercial and


funato RC 3-storey Unknown Unknown No Unknown
apartment blocks

funato Wedding Plaza Unknown 3-storey 3 Unknown No Unknown

funato Nokyo credit co-operative Unknown <4-storey Unknown Unknown No Unknown

funato Kesennuma banking building Unknown Unknown Unknown Unknown No Unknown

Maiya shopping plaza (and Car park steel 24hr access to


funato 5-storey 3 Unknown No
adjacent multi-storey car park) frame car park only

funato Fukutomi Hotel Unknown Unknown Unknown Unknown No Unknown

GNS Science Report 2012/17 78


2012

Inundated
City/town Building Construction Height People saved Signage 24-hour access
storeys

funato Plaza Hotel RC 4-storey 3 22 No External stairs

Staff to bring key


Natori Yuriage Community Centre RC 2-storey 1 43 No and open
buildings

External stairs,
staff to bring key
Natori Yuriage Elementary School RC 3-storey 1 870 No
and open
buildings

External stairs,
staff to bring key
Natori Yuriage Junior High School RC 3-storey 1 823 No
and open
buildings

RC (note lots of
Natori Sendai International Airport 2-storey 1 Unknown No Unknown
external glazing)

Kamaishi Apartment block RC 8-storey 3 Yes Unknown

Local residents
50
provided with a
Kamaishi Government building RC 2-storey Unknown No
key for night-time
access

Kamaishi (beach
Hotel Horaikan Steel frame 4-storey 2 - Yes External stairs
loc. Unosumai)

Public instructed
to break windows
Kesennuma Government building RC 5-storey 2 120 No
to gain access
when closed

GNS Science Report 2012/17 79


2012

Inundated
City/town Building Construction Height People saved Signage 24-hour access
storeys

Security staff to
Kesennuma Prefectural building RC 5-storey 2 200 Yes
let people enter

Kesennuma Central community centre RC 3-storey 2 450 No External stairs

Kesennuma Centre for working young RC 2-storey 2 1 No Unknown

External stairs,
vehicle ramp to
Kesennuma Fish market Steel frame 1-storey 1 1,000 No
car park on
inland side

Kesennuma Park Hotel Unknown Unknown Unknown - No Open 24-hrs

Kesennuma Kahoku newspaper building RC 3-storey 1 71 No Unknown

Minami Kesennuma elementary


Kesennuma Unknown 3-storey 1 Unknown No Unknown
school

Public instructed
to break windows
Kesennuma Jonan Junior High School Unknown Unknown Unknown - No
to gain access
when closed

Kesennuma Ace Port RC 3-storey 1 - No Unknown

Kesennuma Welfare centre RC 3-storey 2 80 No Unknown

Kesennuma Shoes store Unknown Unknown Unknown 5 No Unknown

Kesennuma Yayoi food factory Steel frame 4-storey 4 400 Yes Open 24-hrs

Kesennuma Koyo High School Unknown Unknown Unknown 49 No Unknown

GNS Science Report 2012/17 80


2012

Inundated
City/town Building Construction Height People saved Signage 24-hour access
storeys

Kesennuma Iwaizaki Centre Unknown Unknown Unknown - No Unknown

Kesennuma Ikkeikaku Hotel Unknown Unknown Unknown 50 No Unknown

Totals 37 buildings 5428 people

GNS Science Report 2012/17 81


Principal Location Other Locations

1 Fairway Drive Dunedin Research Centre Wairakei Research Centre National Isotope Centre
Avalon 764 Cumberland Street 114 Karetoto Road 30 Gracefield Road
PO Box 30368 Private Bag 1930 Wairakei PO Box 31312
Lower Hutt Dunedin Private Bag 2000, Taupo Lower Hutt
New Zealand New Zealand New Zealand New Zealand
T +64-4-570 1444 T +64-3-477 4050 T +64-7-374 8211 T +64-4-570 1444
www.gns.cri.nz F +64-4-570 4600 F +64-3-477 5232 F +64-7-374 8199 F +64-4-570 4657

You might also like